Sunteți pe pagina 1din 89

The Pennsylvania State University The Graduate School

DIRECT FIELD-ORIENTED CONTROL OF AN INDUCTION MACHINE USING AN ADAPTIVE ROTOR RESISTANCE ESTIMATOR

A Thesis in Electrical Engineering by David M. Reed

c 2009 David M. Reed

Submitted in Partial Fulllment of the Requirements for the Degree of

Master of Science

August 2009

The thesis of David M. Reed was reviewed and approved by the following:

Heath F. Hofmann Associate Professor of Electrical Engineering Thesis Advisor

Jerey L. Schiano Associate Professor of Electrical Engineering

W. Kenneth Jenkins Professor of Electrical Engineering Head of the Department of Electrical Engineering

Signatures are on le in the Graduate School.

Abstract

When eld oriented-control techniques are applied to induction machines, a DC machine-like torque response may be achieved, making the induction machine a viable candidate for applications previously dominated by other electric machines. Some of these applications, however, demand high performance over a range of operating conditions. This becomes an issue as the eld-oriented techniques needed to achieve the desired performance also have a well-known sensitivity to variations in machine parameters. While methods exist to account for the variation in magnetic parameters due to saturation, compensation for variations in rotor resistance, which varies with temperature and frequency, remains an open research topic. This thesis presents a new adaptive rotor resistance estimator, for which stability and convergence are rigorously proven, along with rotor ux linkage estimators derived using the trapezoidal rule for improved accuracy and stability. These estimators are then incorporated into a direct eld-oriented controller and thoroughly tested on experimental hardware. The resulting controller achieves an acceptable level of performance over a range of operating conditions.

iii

Table of Contents

List of Figures List of Tables List of Commonly Used Symbols Acknowledgments Chapter 1 Field-Oriented Control of Induction Machines 1.1 Basic Drive System and Background . . . . . . . . . 1.2 Field-Oriented Control . . . . . . . . . . . . . . . . . 1.2.1 Basic Principles of Field-Oriented Control . . 1.2.2 Shortcomings of Field Oriented Control . . . . 1.2.3 Rotor Resistance Compensation In Literature 1.3 Thesis Contributions and Organization . . . . . . . . Chapter 2 Induction Machine Model 2.1 Smooth Airgap Dynamic d q 2.2 Observability Analysis . . . . 2.2.1 State Space Model . . 2.2.2 Observability Test . . 2.3 Simulink R Model . . . . . . .

vi viii ix xi

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

1 1 6 6 8 10 13

Model . . . . . . . . . . . . . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

15 15 20 20 21 22

Chapter 3 Controller Design 25 3.1 Direct Field-Oriented Controller . . . . . . . . . . . . . . . . . . . . 25 iv

3.2

3.3

3.4

Rotor Flux Estimator Design . . . . . . . . . . 3.2.1 Derivation using Trapezoidal Integration 3.2.1.1 Voltage-Based Estimator . . . . 3.2.1.2 Current-Based Estimator . . . 3.2.2 Dead-Time Eect and Compensation . . Adaptive Rotor Resistance Estimator . . . . . . 3.3.1 Derivation and Lyapunov Stability Proof 3.3.2 Steady State Analysis . . . . . . . . . . Simulink Simulation . . . . . . . . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

28 30 31 34 36 39 40 43 45

Chapter 4 Implementation and Experimental Verication 4.1 Induction Machine Parameter Estimation . . . . 4.1.1 Parameter Estimation Technique . . . . 4.1.2 Experimental Setup and Results . . . . . 4.2 Dead Time Measurement . . . . . . . . . . . . . 4.3 Controller Experimental Verication . . . . . . 4.3.1 Implementation . . . . . . . . . . . . . . 4.3.2 Experimental Verication . . . . . . . . Chapter 5 Discussion and Future Work 5.1 Controller Performance and Limitations 5.1.1 Dependence on Operating Point . 5.1.2 Flux Estimator Performance . . . 5.2 Suggestions for Future Work . . . . . . . Bibliography

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

50 50 51 53 56 58 58 64

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

69 69 70 71 73 76

List of Figures
1.1 1.2 1.3 1.4 Drive system archetecture. . . . . . . . . . . . . . . . . . . . . . . Rotor temperature increase with respect to time for a 1.5 kW at rated operating conditions [14]. . . . . . . . . . . . . . . . . . . . Park transform vector diagram. . . . . . . . . . . . . . . . . . . . Simulated torque responses for a 1 N-m command with Rr mismatch using direct eld-oriented control with ux estimation for torque regulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Smooth air-gap 2-phase (d q ) induction machine cross-section. Induction machine magnetics model. . . . . . . . . . . . . . . . Simulink block diagram for the current/ux relationships. . . . . Simulink block diagram for the electrical dynamics. . . . . . . . Simulink block diagram for the mechanical dynamics. . . . . . . . . . . . . . . 2 6 7

. . . . . .

10 16 17 23 24 24 26 28 32 35 36 37 38 46 47 47 48 48 49

2.1 2.2 2.3 2.4 2.5 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 3.10 3.11 3.12 3.13

Transformation of electrical variables into the rotor ux linkage reference frame. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Direct Field-Oriented Controller. . . . . . . . . . . . . . . . . . . . Frequency response comparison of integrator approximations. . . . . Pole locus comparison of Forward Euler and Trapezoidal methods. . Inverter IGBT totem pair. . . . . . . . . . . . . . . . . . . . . . . . Inverter waveforms with dead-time eect. . . . . . . . . . . . . . . . Illustration of the rst harmonic of the dead-time voltage. . . . . . Simulation results. . . . . . . . . . . . . . . . . . . . . . . . . . . . Simulink block diagram of resistance estimator and reference frame calculator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Simulink block diagram of voltage-based rotor ux linkage estimator. Simulink block diagram of current-based rotor ux linkage estimator. Simulink block diagram of the direct eld-oriented controller. . . . . Simulink block diagrams of the Modied (left ) and Inverse (right ) Clark transforms. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

vi

3.14 Simulink block diagrams of the Park (left ) and Inverse Park (right ) transforms. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.15 Simulink block diagram of a PI regulator. . . . . . . . . . . . . . . . 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 4.10 4.11 4.12 4.13 4.14 4.15 4.16 4.17 4.18 4.19 5.1 5.2 5.3 5.4 Stator current locus as a function of slip frequency in the stator ux linkage reference frame. . . . . . . . . . . . . . . . . . . . . . Experimental setup. . . . . . . . . . . . . . . . . . . . . . . . . . Simulink block diagram for parameter estimation experiment. . . Comparison of experimental data and tted stator current locus (left ) with normalized error (right ). . . . . . . . . . . . . . . . . . Dead time measurement experimental setup. . . . . . . . . . . . . Dead time measurement for a 20 percent duty cycle. . . . . . . . . IGBT Bridge Conguration. . . . . . . . . . . . . . . . . . . . . . dSpace DS1104 block diagram [30]. . . . . . . . . . . . . . . . . . Simulink block diagram of controller implementation for dSpace code generation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . Simulink block diagram of duty cycle generator. . . . . . . . . . . Simulink block diagram of second order integrator approximation for the voltage-based estimator. . . . . . . . . . . . . . . . . . . . Simulink block diagram of coecient calculation for the second order integrator approximation. . . . . . . . . . . . . . . . . . . . . Simulink block diagram of current-based estimator. . . . . . . . . Simulink block diagrams of the current-based estimator coecients a[k] (left ) and b[k] (right ). . . . . . . . . . . . . . . . . . . . . . . Simulink block diagrams of the current-based estimator coecients c[k] (left ) and d[k] (right ). . . . . . . . . . . . . . . . . . . . . . . Simulink block diagram of the current-based estimator coecient e[k]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Adaptive estimator experimental transient responses. . . . . . . . Experimental response to a torque step from 1 N-m to 3 N-m. . . Comparison of torque steps for mismatched and tuned rotor resistances. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

49 49 51 54 55 56 57 57 58 59 60 61 62 62 63 63 63 63 64 66 68 70 72 73 74

Steady-state rotor resistance estimation at dierent rotor velocities as a function of slip frequency. . . . . . . . . . . . . . . . . . . . . . Open-loop characterization of agreement between the voltage and current-based rotor ux linkage estimators. . . . . . . . . . . . . . . Induction machine model incorporating magnetic saturation eects. Block diagram of the proposed method for turning the resistance estimator ON and OFF. . . . . . . . . . . . . . . . . . . . . . . . . vii

List of Tables
3.1 3.2 4.1 4.2 Parameters used for pole locus comparison. . . . . . . . . . . . . . . Simulation parameters. . . . . . . . . . . . . . . . . . . . . . . . . . Manufacturer machine ratings. . . . . . . . . . . . . . . . . . . . . . Estimated test machine parameters. . . . . . . . . . . . . . . . . . . 35 45 54 56

viii

List of Commonly Used Symbols


s = [sd , sq ] r = [rd , rq ] is = [isd , isq ] ir = [ird , irq ]
T T T T T T

Stator Flux-Linkage Vector. Rotor Flux-Linkage Vector. Stator Current Vector. Rotor Current Vector. Stator Voltage Vector. Rotor Voltage Vector. Stator Resistance. Rotor Resistance. Stator Self-Inductance. Rotor Self-Inductance. Mutual Inductance. Leakage Term. Electromagnetic Torque. Three-Phase Electromagnetic Torque. Load Torque. Number of Poles. Electrical Frequency. Rotor Angular Velocity.

vs = [vsd , vsq ] vr = [vrd , vrq ] Rs Rr Ls Lr M

2 = Ls Lr M 2 e 3ph l P e r re =
P 2 r

Electrical Rotor Angular Velocity.

ix

se = e re H B J= I= 0= 0 1 1 0 1 0 0 1 0 0 0 0

Electrical Slip Frequency. Moment of Inertia. Mechanical Damping. 90 Rotation Matrix. The Identity Matrix. The Zero Matrix.

Acknowledgments
I would like to thank Dr. Heath Hofmann for his guidance throughout this project. I have learned a great deal from him, and with him, in the past year. I would also like to thank Dr. Je Schiano for serving on my committee, and teaching me a great deal about control theory and design. Additionally, I wish to thank Dr. Jack Mitchell, Prof. Mark Wharton, Dr. Je Mayer, Dr. Javier Gomez-Calderon, and all the teachers who have inspired me throughout my education. I have been fortunate to have had so many excellent teachers. Finally, I would like to thank my family and friends for all of their love and support. Because of you, I can safely say that I have come out of grad school with my sanity intact.

xi

Chapter

Field-Oriented Control of Induction Machines


Field-oriented control (also referred to as vector control) is often used in highperformance drive applications. The technique, conceived in the early 1970s, achieves DC machine-like performance from induction machines; and while computationally intensive, the availability of cheap, powerful microprocessors has made eld-oriented control a viable choice for modern high performance drive systems.

1.1

Basic Drive System and Background

The term electric drive generally refers to the power electronics, controller and electrical sensors required to operate an electric machine in specic applications where control over torque, speed and operating points is desired. While electric drives can be used with a variety of electric machines (AC and DC), the focus of this thesis will be the control of induction machines using an electric drive. The basic motor drive system architecture, shown in Figure 1.1, gives us control over the frequency of the AC voltage used to drive the motor.

(dSpace)

va vb vc

Figure 1.1. Drive system archetecture.

It should be noted that, while the focus of this thesis is the control of 3phase induction machines, the control techniques used are based on a two-phase equivalent model which will be discussed in the following chapter. The use of twophase models is common practice as it leads to reduced order models which are easier to work with, and may be used with higher-than-three-phase machines as well. For three-phase machines, the electrical variables (a,b,c) can be transformed into their equivalent two-phase values (d,q,0) using the Clark transform (1.1). xd 2/3 1/3 1/3 xa xa x = 0 3/3 3/3 x = T23 x q b b 1/3 1/3 1/3 x0 xc xc Likewise, the inverse Clark transform is given by 0 1 xd xa 1 xd xd 1 x x = 1/2 3/2 1 x = T = T x 32 23 q b q q xc 1/2 3/2 1 xo xo xo

(1.1)

(1.2)

3 A simplication can be made by taking the zero sequence component (xo ) to be equal to zero. This common assumption, which will be used throughout this thesis, leads to the following expression known as the Modied Clark transform (1.3). xd 1 = 3 xq 3 0 xa 2 3 xb 3 (1.3)

Electric drives can be used in a variety of applications ranging from motion control (servo drives) to vehicle propulsion (traction drives). The increasing aordability of the necessary electronics continues to make electric drives a cost-eective alternative for applications once dominated by mechanical systems. Hydraulic lines and mechanical linkages/cables can be replaced by electric wires and motors [1]. For example, many new cars now employ electric throttle and transmission linkages. While there are practical concerns with such systems (loss of electric power means loss of electrical systems) they also have advantages over their mechanical alternatives. The dynamics of electrical systems are typically much faster than those of mechanical systems, and electrical power is often easier to transport than fuel or hydraulic uid. Also, when designed properly, electrical systems can prove to be more reliable than mechanical systems. Induction machines in particular are renowned for their low maintenance and durability. Recently, the rising cost of energy and environmental concerns has sparked interest in hybrid vehicles such as the Toyota Prius, and electric vehicles such as the Tesla Roadster. The use of electric drives for vehicle propulsion presents an interesting challenge, particularly in the case of an all-electric vehicle such as the Tesla Roadster. Such applications require performance over a very wide speed range and under a variety

4 of operating conditions (loads, temperatures, etc.). Achieving an acceptable level of performance under these circumstances is no easy task. This is especially true when using an induction machine for propulsion. Machine parameters can vary with electrical frequency, ux levels and temperature. These variations tend to detune the electric drives control system and degrade its performance. Electrical frequency aects eciency, resistance, and controller performance. The losses due to hysteresis and eddy currents increase with frequency, and depending on the characteristics of the conductors in a machine, the skin eect can signicantly increase the eective rotor and stator resistances. Controller performance is aected by frequency as a result of necessary approximations made when implementing the controller on the dSpace microprocessor. One concern in particular is the ability to accurately estimate ux linkages over wide speed ranges. However, this issue has been addressed by Jansen, Thompson and Lorenz in [17]. The magnetic parameters, while modeled as linear in our work, are actually nonlinear in nature. This nonlinearity is due to the saturation of magnetic materials in the machine under high magnetic eld strengths. The result is that the inductances, which are modeled as constant under the assumption of linearity, are actually a nonlinear function of magnetic eld strength. There are, however, methods of accounting for this nonlinearity. One approach is to make the mutual inductance between the stator and rotor, a nonlinear function of the air-gap ux linkage magnitude [1]. Naturally, temperature plays a major role in the variation of parameters in an induction machine. The most signicantly aected parameters are the stator and rotor resistances. In particular, variations in the rotor resistance have a more dramatic eect on controller performance. For this reason, many attempts at accounting for this variation have been made [19]-[28]. This variation in rotor

5 resistance is the primary motivation for the work presented in this thesis. The resistance change with temperature, in this case for the rotor, can be described by the following equation [14]:

Rr = Rr,cold (1 + p )

(1.4)

where is the temperature coecient (typically for aluminum, Al =

0.0042 C )

and

p = rotor ambient is the rotor temperature increase. In [14] the authors used a rst order equivalent RC circuit to model the rotor temperature increase: t dt = ploss dt Rt

Ct

(1.5)

where Ct and Rt are the thermal capacitance and resistance respectively, and the subscript t refers to the thermal model. Least-squares was used to estimate the thermal capacitance and resistance from experimental data for a 1.5 kW induction machine running at the manufacturers rated conditions. The resulting thermal time constant, Tt = Rt Ct , was found to be 2490 seconds. Their plot, evaluating the accuracy of the thermal model (1.5) has been provided in Figure 1.2. Inspection of these results yields two important observations as it concerns this thesis. First, using the data in Figure 1.2 and (1.4), we nd that the rotor resistance would have increased approximately 25%. Secondly, inspection of Figure 1.2 and the thermal time constant, Tt = 2490s, reveals that the rotor resistance will change very slowly with respect to time. This will allow us to make an important assumption when deriving our adaptive rotor resistance estimator in a later chapter.

DATA OF THE INVESTIGATED INDUCTION MACHINES

sistance for various iron loss components as a actual iron losses are represented by the best : for the investigated tal curve (p machine).

= 0 0154

the reference temperature rise . he total losses (20) are required in order parameters in (24). A two-dimensional and enables the minimization ers e error of the estimated temperature difFigure 1.2. Rotor temperature with respect for a 1.5 kW at rated with respect increase to the thermal model (1# )to fortime the investigated (1# ) and (32) operating conditions [14]. machine. 1.5-kW induction
Fig. 7. Estimated rotor temperature with respect to the parameter model

btained measuring points in the time do-

1.2

rameter model from the temperature estimated by the thermal Field-Oriented Control model is less than 3 C for all three motors.

The units of the thermal resistor and the thermal capacitor are (33) mixed p.u. and SI units.in Usually, the physical unit of the thermal As mentioned earlier, the interest eld-oriented control stems from the ability to . In this case the motor losses are derived as resistor is of rotor temperature with respect p.u. quantities. Therefore, unit of the thermal resistor in this performance) obtainto a the nearly instantaneous torquethe response (i.e. DC machine-like ed in Figs. 79 for the investigated mo- paper is only . The physical meaning of the thermal resistor in well induction This performance is the achieved by exploiting the natural deand as as the machines. mal parameters can be explained with the help of steady-state temperature of (33) (34)

coupling which results when the characteristic equations for an induction machine are written in the rotor-ux reference frame.
(35) If the p.u. total losses are, e.g., 5% of the reference power (5), the steady-state rotor temperature is 5% of . The physical unit of a

chines are summarized in Table II. The f the temperature estimated by the pa-

1.2.1

Basic Principles of Field-Oriented Control

use limited to: Penn State University. Downloaded on June 14, 2009 at 15:49 from IEEE Xplore. Restrictions apply.

A fundamental feature of eld-oriented control is the transformation of electrical variables into a rotating synchronous reference frame [Figure 1.3]. This transformation is often referred to as the Park transform (1.6), named for R.H. Park who published the rst papers in 1929, detailing the application of reference frame theory to the analysis of AC machines [13].

syn q syn syn q syn x d syn d

Figure 1.3. Park transform vector diagram.

syn

syn xd cos (syn ) sin (syn ) xd J = = = e syn x syn sin (syn ) cos (syn ) xq xq

(1.6)

For completeness, the inverse Park transform is given by syn xd cos (syn ) sin (syn ) xd J x= = = e syn x sin (syn ) cos (syn ) xsyn xq q

syn

(1.7)

It has been shown [1] that the torque expressions for various AC machines simplify when expressed in certain synchronous reference frames. For this reason Park transforms are used in eld-oriented control, despite the added complexity. Once transformed into a synchronous reference frame, standard control techniques can be applied. The output is then transformed back into a stationary reference frame using the inverse Park transform (1.7) [1]. A common analogy used to describe the operation of eld-oriented AC drives,

8 is that of a separately excited DC motor drive [2]. The general torque expression for a DC machine is given by P af (if ) ia , 2

e =

(1.8)

where P is the number of poles, af (if ) is the armature ux linkage expression which is a function of the eld winding current, if , in a separately excited DC machine, and ia is the armature current. Inspection of (1.8) reveals that if can be used to regulate the armature ux linkage, and ia can be used to regulate torque. This decoupling is inherent in DC machines because the elds produced by ia and if are orthogonal by design. However, it will later be shown that this decoupling and a similar torque expression can be achieved for induction machines by transforming the electrical variables into the rotor ux linkage reference frame. In general, there are two types of eld-oriented controllers for induction machines which are based on the rotor ux linkage reference frame, indirect (or feedforward) and direct (or feedback). The latter method will serve as the basis for the controller proposed in this thesis.

1.2.2

Shortcomings of Field Oriented Control

Despite the high performance which can be achieved using eld-oriented control, there are shortcomings as well. There are two primary methods of estimating ux linkages for use in control algorithms, which will be discussed in detail later. If a ux linkage estimator is used in the control algorithm, performance at either low frequencies or high frequencies will suer depending on the type of estimator used. These estimators can also suer from parameter variations, and if speed-sensorless control is being performed, the loss of observability at DC excitation [16] will likely

9 need to be addressed [18]. For simplicity in our work, we allowed ourselves the use of a position encoder, from which rotor speed was derived. While direct eld-oriented controllers can be implemented by directly measuring the magnitude and position of the aig-gap ux in the machine using either pickup coil or hall-eect sensors, this method is not favored as it requires modications to the machine to place the sensors in the air-gap, and will still be prone to error. Therefore, direct eld-oriented controllers are often implemented using estimators to determine the rotor ux linkage. Both the indirect and direct methods tend to lack robustness, having similar sensitivities to variations in machine parameters [15]. In fact, when a ux estimator is used with the direct method, the parameter variation of concern for both methods is that of the rotor resistance, which is a function of temperature and frequency. This is because a change in the rotor resistance, which may exceed 50% [2], causes a change in what is referred to as the rotor time constant, Tr = Lr /Rr . The accuracy of this parameter is critical for the slip frequency computation performed when using the indirect method. Likewise, it is of importance in the ux estimators commonly used to implement the direct method. A mismatch between the actual rotor resistance and the value used in a eldoriented controller can lead to signicant steady-state errors in torque regulation if the mismatch is severe enough. To demonstrate the potential severity of such a mismatch, Simulink simulations were performed for a 1 N-m torque command using a direct eld-oriented controller with 50% variations in the value of Rr [Figure 1.4]. Inspection of the torque responses in Figure 1.4 reveals that while the steady-state error due to rotor resistance mismatches can be rather large, there are also oscillations in the response that indicate that the stability of the system could be degraded as well. For these reasons, a great deal of research has been

10

1.6 1.4 1.2

Rr = + 50%

3ph [N-m]

1 0.8 0.6 0.4 0.2 0 0

Rr = 0%

R = - 50%
r

0.5

1.5

2.5

Time [Seconds]
Figure 1.4. Simulated torque responses for a 1 N-m command with Rr mismatch using direct eld-oriented control with ux estimation for torque regulation.

done to nd a practical method for updating the value of the rotor resistance in eld-oriented control algorithms.

1.2.3

Rotor Resistance Compensation In Literature

Addressing rotor resistance variation in induction machines has been a popular research topic for the past two decades. Despite the variety of approaches [19][28], no one method emerged as a widely accepted technique. The Model Reference Adaptive System (MRAS) schemes proposed in [21, 22] are based on the indirect eld-oriented control technique. However, the scheme proposed in [21] does not include a stability proof, and the controller in [22] has yet to be tested on an experimental setup. Another MRAS scheme proposed in [23] is based on the direct eld-oriented control technique. However, their proposed rotor time constant

11 update (1.9) will result in a singularity as the slip frequency goes to zero, sl 0. r r Tr sl

Tr =

(1.9)

Other adaptive techniques have been proposed as well [24]-[28]. However, the controller proposed in [24] uses adaptation laws derived from a dynamic model of the induction machine which is based on the stator current reference frame. Since the electrical variables used in the adaptive estimator are constant in such a reference frame, the system lacks persistency of excitation [10] under steadystate operating conditions, and so it must rely on transients to tune the resistance estimation. A non-linear adaptive sliding mode controller was derived in [25]; however, it has not been tested on hardware. The other adaptive estimation papers [26]-[28] present the evolution of a speed-sensorless direct eld-oriented controller with adaptive rotor time constant estimation. Their rst attempt [26] estimated r , as being proportional to their stator resistance adaptation: the rotor resistance, R d d Rr = Rsrn R s = 1 (eids ids + eiqs iqs ) dt dt

(1.10)

where eids = ids ids , eiqs = iqs iqs , 1 > 0 is a control gain, and Rsrn is the ratio of the nominal values of the stator and rotor resistance. This method is awed however, as the stator and rotor resistances will vary quite dierently for a number of reasons. First, the stator windings consist of copper wire while the rotor cage consists of aluminum bars. Secondly, the rotor and stator are separated by an air-gap which makes it unlikely that the stator and rotor temperatures will be the same. Therefore, the relation between the stator resistance and rotor resistance cannot be described by a simple constant. The next evolution of the controller [27] proposes a new rotor time constant

12 estimator given by: d dr M qr M (1/ r ) = 2 /Lr eids ( ids ) + eiqs ( iqs ) dt

(1.11)

where eids = ids ids , eiqs = iqs iqs , and 2 is an arbitrary positive gain. The authors direct readers to a prior conference paper [29] in which the estimator was proposed and stability veried using the Lyapunov stability criteria. Their proof however, only yields a negative semi-denite result, as stated by the authors. This result only proves that the estimator is stable in the sense of Lyapunov, it does not guarantee that the rotor time constant estimation will converge to the true value (i.e. er 0 as t ). Later evolutions of this estimator [28] are all derived from the original adaptation law (1.11) by rewriting it in the electrical reference frame and in terms of rotor resistance. Therefore, the convergence of these later incarnations is also unproven. Other techniques for updating the rotor time constant have been proposed as well [19, 20]. The method proposed in [19], uses a modied switching scheme to extend the zero crossing of a particular phase current. This forces the current in one of the phases to be zero, simplifying the dynamic equations for the induction machine and allowing a direct calculation of the rotor time constant using known and measurable parameters. However, the accurate convergence of this estimation technique is unproven and would likely be dicult to prove. Additionally, the modied switching technique used to implement the estimator is for Current Regulated Pulse-Width-Modulated (CRPWM) converters and it is unclear how it could aect the performance of the eld-oriented controller. Lastly, the authors only provide experimental results for the implementation of the switching scheme and not the rotor time constant estimation.

13

1.3

Thesis Contributions and Organization

Sensitivity to parameter variations is the leading drawback to the use of eldoriented control techniques with induction machines. Despite the large research eort [19]-[28], the problem of compensating for rotor resistance variations remains an open topic. While previous attempts have had some success, very few suciently explored the practical limitations of their estimators. Slip frequency, integrator approximations and unmodeled eects such as dead-time and saturation can signicantly aect the performance of rotor resistance estimators. This thesis addresses the rotor resistance variation using a new adaptive rotor resistance estimator for which convergence is rigorously proven using Lyapunovs stability criterion, Barbalats lemma, and two-time-scale theory [11]. Rotor ux linkage estimators are derived using trapezoidal integration to provide accuracy and stability over a suciently wide speed range. The resulting rotor ux and resistance estimators are incorporated into a direct eld-oriented controller and tested over a range of rotor speed and slip frequency. To summarize, the contributions of this thesis are the following: Rotor Resistance Estimator: A new rotor resistance estimator is presented along with a thorough analysis of its stability and convergence. Rotor Flux Linkage Estimators: Derivations of the standard voltagebased and current-based rotor ux linkage estimators using trapezoidal integration are presented. Thorough Evaluation of Estimator Limitations: The proposed rotor resistance estimator and ux linkage estimators are thoroughly tested on experimental hardware to reveal limitations.

14 The organization of this thesis is as follows. The standard smooth airgap twophase induction machine model is presented in Chapter 2 in order to familiarize the reader with the model and assumptions which will serve as the basis of our control design. The following sections of Chapter 2 present an observability analysis of the electrical variables necessary to our control design, followed by the implementation of the induction machine model in Simulink. The theoretical details of our control design are presented in Chapter 3. We begin by introducing the basic direct eld-oriented control theory followed by the derivation of our rotor ux linkage estimators using trapezoidal integration. A subsection discussing the dead-time eect and a method for its compensation is included as well. The next section details the derivation and stability proof of our rotor resistance estimator, and includes a steady-state analysis as well. The chapter concludes with a Simulink implementation and simulation results for an idealized implementation of the resulting controller. Chapter 4 begins with a description of our induction machine parameter estimation technique along with the experimental setup and resulting machine parameters. We then discuss our experimental measurement of the dead time, and conclude with the experimental verication of our direct eld-oriented controller and adaptive rotor resistance estimator. Finally, Chapter 5 discusses the performance of our controller over a range of rotor speed and slip frequencies. The limitations which were predicted and discovered are summarized as well. The chapter concludes with recommendations for future work and suggestions for a practical implementation of the resulting controller.

Chapter

Induction Machine Model


There are a variety of ways to model the operation of an induction machine. Naturally, the design of high performance drive systems requires the use of a dynamical model for the induction machine. The model which will be used for the remainder of this document is referred to as the Smooth Airgap Dynamic d q Model. This model will be presented along with an observability analysis and Simulink implementation of the model.

2.1

Smooth Airgap Dynamic d q Model

To simplify analysis, we have modeled the induction machine as having a smooth airgap, and so slot harmonics can be neglected. Additionally, the following assumptions have been made: Linear magnetics model (Neglect saturation and hysteresis). Lossless core (Neglect hysteresis and eddy current losses). Balanced construction with sinusoidally distributed magneto motive force (mmf).

16 1:1 eective turns ratio. These assumptions are typical and appropriate under normal operating conditions. A cross-sectional depiction of the smooth air-gap d q model is provided in Figure 2.1.

Figure 2.1. Smooth air-gap 2-phase (d q ) induction machine cross-section.

As mentioned earlier, the dynamic d q (two-phase) model of the induction machine serves as the basis for the eld-oriented control techniques which will be used in the proposed controller. This model describes the dynamical behavior of a 3-phase induction machine as an equivalent mmf 2-phase machine. The resulting 2-phase (d q ) parameters (voltage, current, etc.) can then be transformed into equivalent 3-phase values using the inverse Clark transform (1.2). Note that under balanced operation the zero sequence parameter xo is equal to zero, and so can

17 be neglected. The order reduction achieved by using the d q model can now be appriciated as the charateristic equations for a two-phase equivalent induction machine model [Figure 2.2] are presented.

ls

lr

re r

Figure 2.2. Induction machine magnetics model.

The stator and rotor ux linkages (designated by ) in an arbitrary reference frame (designated by the superscript x) are given by the following relationships:

sx = Ls isx + M ir x , r x = M isx + Lr ir x ,

(2.1)

(2.2)

where Ls = Lls + M and Lr = Llr + M , are the stator and rotor self-inductances, isx and ir x are the stator and rotor currents, M is the mutual inductance, Lls and Llr are leakage inductances. The stator voltage in the stationary reference frame can be found using Kirchos Voltage Law (KVL) and Faradays law: ds dt

vs = Rs is +

(2.3)

18 Similarly, the rotor voltage expressed in the rotor electrical reference frame (denoted by the superscript re), is given by the same relationship as above: dr re dt

vr re = Rr ir re +

(2.4)

This relationship can then be expressed in the stationary reference frame by rewriting the previous equation as follows:

vr = eJre vr re = eJre Rr ir re + dr re dt d Jre e r dt dr dt (2.5)

= Rr eJre eJre ir + eJre = Rr ir + eJre

Jre eJre r + eJre

Finally, we arrive at our desired expression for the rotor electrical dynamics: dr =0 dt

vr = Rr ir re Jr +

(2.6)

It should be noted that in our analysis, we have modeled the induction machine as having a squirrel-cage rotor winding and so the rotor voltage is equal to zero (the terminals are shorted together). Additionally, the machine used in our experimental testing was in fact a squirrel-cage type induction machine. To derive the torque expression for an induction machine, we will use the concept of co-energy [1, 3, 4]. It can be shown that the expression for co-energy is given by 1 Wc = Ls is 2
2

1 + Lr ir re 2

+ isT M (re ) ir re

(2.7)

19 where M (re ) = M eJre . This expression may look familiar as it is strikingly similar to the equation for energy stored in a transformer. The electromagnetic torque of a three-phase machine is then given by 3 Wc 3 Wc re = 2 r 2 re r

3ph =

(2.8)

where

P re = . r 2 Substituting (2.7) into (2.8) yields the following torque expression (2.9). 3P T is M JeJre ir re 4

3ph =

(2.9)

While the torque expression above (2.9) is valid, it is not particularly useful due to its dependence on the rotor electrical angular displacement (re ). This dependence can be eliminated however, by recognizing that ir re = eJre ir , and so we obtain our desired result: 3P M i s T Ji r 4

3ph =

(2.10)

Lastly, the mechanical dynamics are given by dr 1 = [(3ph l ) Br ] dt H

(2.11)

where H is the combined moment of inertia of the rotor and load, B is the mechanical damping on the rotor, and l is the load torque applied to the rotor.

20

2.2

Observability Analysis

As mentioned earlier, direct eld-oriented controllers are typicially implemented using a rotor ux linkage estimator. The estimator is needed to transform the electrical variables into and out of the rotor ux linkage reference frame. Additionally, an estimation of the rotor ux linkage magnitude is computed as well, which is then used in the control algorithm. Given the importance of the rotor ux linkage estimator, an observability analysis is well justied.

2.2.1

State Space Model

The observability analysis presented in this thesis will be limited to the electrical dynamics of the induction machine. While the rotor velocity, whose dynamics are given by (2.11), will be present in the state equations, we will argue that the time scale associated with them is considerably slower than that of the electrical dynamics. Therefore, the rotor velocity will be treated as a constant in our analysis, since in reality it is changing much more slowly than the electrical variables. Additionally, we will be measuring the rotor velocity with an optical encoder in our experiments, and so an instantaneous measurement will always be available. The same argument will be used in the following chapter to derive ux estimators using trapezoidal integration. By making this assumption we not only reduce the order of the state space equations, but also obtain a linear model. d x dt
e

= Ax e + Bvs e is e = C x

(2.12)

The state space equations in the electrical reference frame (denoted by the superscript e) for the smooth airgap d q model of an induction machine [16] can

21 be derived using the equations presented in the previous section. The resulting vectors and matrices are given as follows:
e

(2.13)

is = r e

2 M Rr (Rs L2 r + Rr M ) I J I J e re 2 Lr 2 Lr A= Rr Rr M I I (e re ) J Lr Lr Lr I B = 2 0

(2.14)

(2.15)

C= I 0

(2.16)

where 2 = Ls Lr M 2 . The electrical reference frame is used so that the electrical frequency (e ) appears in the equations. This way, the dependency of the systems observability on electrical frequency, if any, can be analyzed.

2.2.2

Observability Test

To verify the observability of the electrical variables (ise and r e ), we will use the standard test from linear systems theory [7, 8]. The observability matrix can then be formed as follows:

C CA Q= CA2 3 CA

(2.17)

22 Even with the reduction in model order, the result is a rather large and complex 8by-4 matrix. However, from linear algebra we know that the rank of a matrix (Q), which is the maximum number of linearly independent columns, must be less-thanor-equal-to the minimum dimension. Since the matrix Q has a minimum dimension of 4, the rank of the matrix must be less-than-or-equal-to 4. Additionally, the rank of a matrix is also equivalent to the number of linearly independent rows [12]. In our analysis, we seek to prove that the electrical variables are completely observable by showing that the observability matrix has full rank (i.e. (Q) = 4) for all values of e and re . Fortunately, this can be proven by showing that the rst four rows of the observability matrix are linearly independent. These rows are given by 0 Rr I re J Lr (2.18)

Q14

I C = = 2 (Rs L2 M r + Rr M ) CA I e J 2 Lr 2

Inspection of (2.18) reveals that the only way the bottom two rows could be a linear combination of the top two rows is when we and wre both equal zero, along with either M or Rr . While the rst condition (we = wre = 0) is not only possible, but a very likely operating condition, the second condition (M or Rr = 0) is physically impossible. Therefore, the matrix Q has full rank and so the electrical variables ise and r e are completely observable for all non-zero values of we and wre .

2.3

Simulink R Model

Simulations were performed in order to test our controllers stability and performance before implementation with hardware. The software package Simulink was

23 chosen for its familiarity, ease of use and its simple interface with dSpace. However, before a controller could be simulated, we rst needed to implement our induction machine model in Simulink. Our simulink model of an induction machine consists of three layers. The bottom layer [Figure 2.3] implements the current/ux linkage relationships given by isx = ir x = Lr x M x 2 r 2 s Ls x M x 2 s 2 r (2.19) (2.20)

1 lambda _sd Lr/sig^2 Lr/sig^2 Add

1 i_sd

2 lambda _sq

2 i_sq M/sig^2 M/sig^2

M/sig^2 3 i_rd

3 lambda _rd

M/sig^1

Ls/sig^2 Add 2 4 lambda _rq Ls/sig^2 4 i_rq

Figure 2.3. Simulink block diagram for the current/ux relationships.

These relationships can be derived from the ux linkage expressions (2.1 2.2) presented earlier. The electrical dynamics, which are given by equation (2.3) and (2.6), are implemented in the next layer [Figure 2.4]. The nal layer [Figure 2.5] implements the mechanical dynamics (2.11 2.10) and performs the necessary Clark transforms.

24

Rs1 Rs Rs Rs

1 Vsd 2 Vsq 1 s Integrator 3

1 s Integrator

lambda_sd

i_sd

1 i_sd 2 i_sq 3 i_rd 4 i_rq

lambda_sq

i_sq

lambda_rd

i_rd

3 w_re Product

-1 J

1 s Integrator 1
lambda_rq i_rq

Current / Flux Relations Rr 1 Product 1 J1 1 s Integrator 2 Rr1 Rr Rr

Figure 2.4. Simulink block diagram for the electrical dynamics.

a i_sd Vsd 1 v_ a i_sq 2 v_ b 3 v_ c abc -dq 0 Conversion c 0 Terminator b q Vsq i_rd Product 3 *P*M /4 w_re Add i_rq Product 1 IM _ dq _ model 4 Load _Torque w_re 3 *P *M /4 tau_3ph 2 -3 Phase Conversion q c a d b d

1 i_a 2 i_b 3 i_c

1 /H Add 1 1 /H

1 s Integrator

w_r

4 Rotor _ Speed

B B P/2 P /2

Figure 2.5. Simulink block diagram for the mechanical dynamics.

Chapter

Controller Design
Until now, we have kept the discussion of our control algorithm very general, gradually building upon the basic concepts. In this chapter however, we will begin presenting the detailed derivation and design of our controller. The direct eldoriented control technique will be rst be discussed along with the derivation of our rotor ux linkage estimators. Afterwards, the derivation and stability proof of our adaptive rotor resistance estimator will be presented as well as simulations verifying the stability and performance of the resulting controller.

3.1

Direct Field-Oriented Controller

As mentioned earlier in chapter one, the direct eld-oriented control technique which we are using is based on the transformation of electrical variables into and out of the rotor ux linkage reference frame. The direct axis of this reference frame is aligned with the rotor ux vector, as shown in Figure 3.1. The torque expression in the rotor ux linkage reference frame can be derived using the equations from chapter two as follows:

26

syn

syn r r d r syn

Figure 3.1. Transformation of electrical variables into the rotor ux linkage reference frame.

3ph =

3P M isT Jir 4 3P M r = is M isT J 4 Lr Lr M i T Jr M isT Jis Lr s M i T Jr Lr s M (rd isq rq isd ) Lr

3P 4 3P = 4 3P = 4 =

(3.1)

Taking the electrical variables to be in the rotor ux linkage reference frame, the expression above (3.1) can be rewritten as follows: 3P M r r ||r || cos (r syn ) i sq ||r || sin (r syn ) isd 4 Lr 3P M r ||r ||i = sq 4 Lr

3ph =

(3.2)

27 Inspection of (3.2) reveals promising similarities to the torque expression for a separately excited DC machine (1.8) given in chapter one. By transforming electrical variables into the rotor ux linkage reference frame, we can regulate the torque of an induction machine by regulating the rotor ux linkage magnitude,
r ||r ||, and quadrature stator current, i sq .

Conceptually, inspection of Figure 3.1 reveals that, since the direct stator current vector will be aligned with the rotor ux linkage vector in the rotor ux linkage reference frame, it may be used to regulate the rotor ux linkage magnitude. This can be shown analytically as follows: d r = Rr ir + re Jr dt r M = Rr is Lr Lr = re J

+ re Jr (3.3)
T

Rr M Rr I r + is Lr Lr

Noting that in the rotor ux linkage reference frame r = ||r || 0 the following expression d Rr Rr M r ||r || = ||r || + i dt Lr Lr sd

we arrive at

(3.4)

r Therefore, isd can be used as a control input to regulate and shape the dynamics
r of the rotor ux linkage magnitude, while i sq can be used to regulate torque for a

give rotor ux linkage magnitude [Figure 3.2]. One challenge which remains to be addressed is the estimation of the rotor ux linkage. As it has been shown, the rotor ux linkage is needed to perform the necessary reference frame transforms. Since it is impossible to directly measure

28

a b

1
3P M 4 Lr

r r r

Figure 3.2. Direct Field-Oriented Controller.

the rotor ux linkage, estimators must be used to compute it based on electrical parameters which are known or can be measured.

3.2

Rotor Flux Estimator Design

There are two primary methods of estimating the rotor ux linkage in an induction machine. The rst method, sometimes referred to as the stator voltage-based estimator, directly solves for the stator ux linkage by integrating the stator voltage minus the resistive drop (3.5). The stator ux linkage can then be used along with the stator current to calculate the rotor ux linkage (3.6). s =

vs Rs is dt

(3.5)

Lr 2 s i s r = M M

(3.6)

In practical implementations, this pure integration is troublesome as a DC oset error can build up over time, leading to instability. Therefore, a decay constant K is often added to the stator ux linkage estimator (3.7), implementing what is

29 sometimes referred to as a leaky integrator. The value of this decay constant can be adjusted to eliminate the DC drift problems, unfortunately it will also make the estimator unreliable at low frequencies [1]. Therefore, the value should be chosen such that it is large enough to eliminate DC drift, but small enough to allow reliable operation at as low a frequency as possible. s = vs Rs is K s dt

(3.7)

The second method of calculating the rotor ux linkage, often referred to as the current-based estimator, can be implemented by integrating the equation for the rotor electrical dynamics given by (3.3). The resulting expression has been provided below. r = re J Rr M Rr I r + is dt Lr Lr (3.8)

It can be shown that the poles of this estimator are located at Rr j re Lr

p1,2 =

(3.9)

by viewing the estimator as a state-space equation where the system matrix is given by A = re J Rr I Lr (3.10)

Therefore, the current-based estimator is naturally stable and so it tends to yield higher accuracy estimations compared to the stator voltage-based estimator at low frequencies. However, it will be shown in the following section that discretetime implementations of this estimator may suer from accuracy and ultimately stability issues at high speeds depending on the integration method used.

30

3.2.1

Derivation using Trapezoidal Integration

Before presenting the derivation of the estimators using trapezoidal integration, we will rst present the Forward Euler implementation of the current-based estimator along with a stability analysis to motivate the use of trapezoidal integration. As we did in the observability analysis, we will again make the assumption that the mechanical dynamics associated with the rotor are changing much more slowly than the electrical dynamics, and therefore re will be treated as a constant in our analysis. Additionally, the rotor velocity will be measured and so an instantaneous value is available. The forward Euler implementation of the current-based estimator can be derived as follows: r [k + 1] r [k ] = Ts Rr M Rr I r [k ] + i s [k ] Lr Lr

re J

Rr Rr M r [k + 1] = r [k ] + Ts re J i s [k ] I r [k ] + Ts Lr Lr r [k + 1] = 1 Ts Rr Lr Rr M is [k ] I + Ts re J r [k ] + Ts Lr (3.11)

where k is the discrete-time index and Ts is the sampling period. To analyze the stability of this estimator we note that the equation is in a state-space form where the system matrix is given by Rr Lr

A=

1 Ts

I + Ts re J

(3.12)

inspection of which reveals that the eigenvalues (or poles) of the estimator are

31 located at p1,2 = 1 Ts Rr Lr jTs re (3.13)

Therefore, the estimator will only be stable for small values of re , eventually becoming unstable at high speeds when the poles move outside the unit circle, which is the stability boundary for discrete-time systems. It can be shown that this will occur when re =
r 1 1 Ts R Lr

Ts

(3.14)

For this reason, and accuracy concerns, the decision was made to use rotor ux linkage estimators derived using trapezoidal integration.

3.2.1.1

Voltage-Based Estimator

Since the voltage-based estimator (3.5) is only marginally stable due to the pure integration theoretically required to solve for the stator ux linkage, stable approximations to ideal integrators (3.7) are typically used in practice. While the use of a rst order approximation or leaky integrator is typically sucient, our adaptive rotor resistance estimation algorithm necessitates the use of a higher order approximation. This is due in part to sensitivity of our adaptive algorithm to phase errors, as well as the need to completely null DC osets. 1 In the s-domain, an ideal integrator is given by , the rst order approximation s 1 . Our proposed approximation is given by of which is given by s+K Gint (s) = s 2 + 2n + n (3.15)

s2

where is the damping coecient and n is the natural frequency which gives the corner frequency of the lter, much like K in the rst order approximation. A

32 Bode plot has been provided [Figure 3.3] which compares the frequency response of an ideal integrator to the rst and second order approximations with corner frequencies of 1 rad/sec. Inspection of Figure 3.3 reveals that the phase response
40 Magnitude [dB] 20 0 -20 -40 -1 10 100 Phase [deg] 50 0 -50 -100 -1 10 10
0

Ideal First Order (K = 1) Second Order ( = 1, = 0.2)


n

10

10

10

10

10

Frequency [rad/sec]
Figure 3.3. Frequency response comparison of integrator approximations.

of the second order approximation is capable of reaching the ideal shift of 90 degrees before the rst order approximation. To obtain a trapezoidal form of (3.15) for implementation, we rst need to derive state space equations which describe the transfer function. This can be done by splitting the transfer function as follows: Y ( s) P ( s) P (s) 1 = [s] 2 2 U (s) s + 2n + n

Gint (s) =

(3.16)

33 The following state equations can then be derived:

x 1 = x2
2 x 2 = u n x1 2n x2

(3.17) (3.18) (3.19)

y = x2

Next, the trapezoidal rule is applied: Ts {x2 [k + 1] + x2 [k ]} 2 Ts 2 u[k + 1] n x1 [k + 1] 2n x2 [k + 1] 2 (3.21) (3.22)

x1 [k + 1] = x1 [k ] + x2 [k + 1] = x2 [k ] + +

(3.20)

Ts 2 u[k ] n x1 [k ] 2n x2 [k ] 2

y [k ] = x2 [k ]

By assuming u[k + 1] = u[k ] the following equations can be found after some substitutions and rearranging: 1 1 + 2d + d2
2 Ts2 2 Ts u[k ] + 1 + 2d + d2 n 2 2

x1 [k + 1] =

x1 [k ] + Ts x2 [k ] , (3.23)

x2 [k + 1] =

1 2 Ts x1 [k ] + 1 2d d2 x2 [k ] , Ts u[k ] n 1 + 2d + d2

(3.24) (3.25)

y [k ] = x2 [k ],

Ts . Provided the trapezoidal rule was applied correctly, the stability 2 of the lter is assured so long as n and are appropriately chosen. This is because where d = n trapezoidal integration, which is the basis of the bilinear transform, preserves the

34 stability of continuous-time systems by mapping left-half-plane poles to locations inside the unit circle [9]. The continuous-time pole locations are given by

p1,2 = n jn

1 2

(3.26)

Therefore, our second order approximation will be stable so long as n and have positive real values. To apply this integrator approximation to our voltage-based ux estimator we simply let y = sx and u = vsx Rs isx . 3.2.1.2 Current-Based Estimator

The current-based rotor ux linkage estimator can be derived using trapezoidal integration as follows: Ts r [k + 1] = r [k ] + 2 Ts Tr 1 M 1 I r [k + 1] + re J I r [k ] + 2 is [k ] Tr Tr Tr re J M 1 I r [k ] + 2 is [k ] Tr Tr

re J

2+

I Ts re J r [k +1] = 2r [k ]+ Ts 2+ 2+
Ts Tr Ts Tr 2

r [k + 1] =

I + Ts re J 2 + (Ts re )2

Ts Tr

M Ts I + Ts re J r [k ] + 2 i s [k ] Tr (3.27)

r [k + 1] = Ar [k ] + Bis [k ], where 4 (Ts re )2 A= 2+


Ts Tr 2 Ts Tr

I + 4Ts re J (3.28)

+ (Ts re )2
s re I + 2 M TT J r 2

B=

Ts 2M 2+ Tr

Ts Tr 2

(3.29)

2+

Ts Tr

+ (Ts re )2

35 Note that the rotor time constant, Tr = Lr /Rr , was used to slightly simplify the equations. Again, the eigenvalues of the system matrix A are given by 4 (Ts re )2 p1,2 = 2+
Ts Tr 2 Ts Tr 2

j 2+

4Ts re
Ts Tr 2

(3.30)

+ (Ts re )2

+ (Ts re )2

For comparison, the eigenvalues/poles of the system matrix were plotted as a fuction of re , in Matlab, for both the Forward Euler and Trapezoidal implementations using typical parameters [Table 3.1].

0.08 0.06 0.04 0.02 0 -0.02 -0.04 -0.06 -0.08 0.992 0.994 0.996 0.998 1

Trapezoidal Forward Euler Unit-Circle

Imaginary

1.002

1.004

Real

Figure 3.4. Pole locus comparison of Forward Euler and Trapezoidal methods.

Rr [ohms] 1.5

Lr [mH] 113

Ts [msec] 0.1

re min [rad/sec] 0

re max [rad/sec] 1000

Table 3.1. Parameters used for pole locus comparison.

36

3.2.2

Dead-Time Eect and Compensation

While the inverter is largely modeled as exhibiting ideal operation, a particular non-ideal eect which we will account for is known as dead or blanking time. The dead-time, td , is a time delay between the opening of one switch (IGBT turning OFF) and closing of the other (IGBT turning ON) in the totem pair [Figure 3.5], inserted by the gate drive circuitry to prevent short circuiting during switching transitions. This time delay is needed as it takes the power electronic transistors a nite amount of time to transition between ON and OFF states.

D high
BUS

vge, high ix(t) vxn(t)

BUS

BUS

D low vge, low


Figure 3.5. Inverter IGBT totem pair.

When driving inductive loads, such as an electric motor, the output current ix (t) will continue owing during the dead-time interval when both IGBTs are OFF. As a result, this current will bias either Dhigh or Dlow ON depending on its direction. This essentially connects either the positive or negative bus to the output, vxn (t). The eect of dead-time on the output voltage of a single phase of the inverter is depicted in Figure 3.6. It should be noted that the neutral point, n, created by splitting the bus voltage across two capacitors is for conceptual

37

ge, high

ON OFF

ON

ge, low
OFF

ON OFF

BUS

xn d d x s s

BUS

Figure 3.6. Inverter waveforms with dead-time eect.

illustration only. The neutral point is generally not needed in practice and so a single or parallel combination of capacitors are typically used. Assuming that the current doesnt change direction during a given switching period, the average value of the voltage due to the dead-time eect is given as 2td VBU S Ts 2 (3.31)

vdead (t)

= sgn(ix (t))

= sgn(ix (t))td fs VBU S ,

where fs is the switching frequency of the power electronics. Furthermore, it can be shown [1] that the average voltage at the output can be written as the summation

38 of the desired average voltage and the average dead-time voltage:

vxn (t)

= (Dx 0.5)VBU S sgn(ix (t))td fs VBU S = vdesired (t) + vdead (t)

(3.32) (3.33)

where D = tON /Ts is the duty cycle for a particular inverter phase. When viewed on a longer timescale, the average dead-time voltage appears as a squarewave which is 180 degrees out-of-phase with the current for a particular phase [Figure 3.7]. Assuming the current waveforms are sinusoidal, our proposed method of compensating for the dead-time voltage is to compute its rst harmonic and add it to our commanded voltages before computing the rotor ux linkage using the voltage-based estimator. This assumption is valid as our controller regulates the direct and quadrature currents in order to regulate torque.
x dead

Figure 3.7. Illustration of the rst harmonic of the dead-time voltage.

The rst harmonic of the dead-time voltage can be computed as follows: ix 4 td fs VBU S ||ix ||

v dead

(3.34)

The decision to use the rst harmonic was motivated by a concern that the higher order harmonics contained in the squarewave approximation of the dead-time voltage could adversely aect the performance of our controller.

39

3.3

Adaptive Rotor Resistance Estimator

The rotor ux linkage estimators designed in the previous section are an important component in our adaptive rotor resistance estimator. For this reason, a great deal of eort was spent designing them. The nal controller will use the current-based rotor ux linkage estimator, which will receive updated resistance values from the adaptive estimator during acceptable operating conditions. The criteria for these operating conditions will be discussed in greater detail later; however, an obvious criteria can be explained now. Since both rotor ux linkage estimators will be needed to estimate the rotor resistance, the estimation must be performed in a frequency range where both estimators are reliable. As mentioned earlier, the need to approximate an ideal integrator using a stable transfer function makes the voltage-based estimator unreliable at low frequencies near wn . Additionally, while the use of trapezoidal integration greatly improves the accuracy and operating range of the current-based estimator, its reliability will still suer at high frequencies due to hardware limitations such as rounding errors. Therefore, the rotor resistance must be tuned in a midband where both estimators are working accurately. This section will present the derivation of our rotor resistance estimator along with a stability proof using the Lyapunov stability criterion and Barbalats lemma [5, 6]. A quasi-steady-state analysis will also be presented which will prove the asymptotic convergence of the rotor resistance error and reveal some fundamental limitations of the estimator which will help us determine acceptable operating conditions for tuning the rotor resistance.

40

3.3.1

Derivation and Lyapunov Stability Proof

Assuming that the magnetic parameters are constant and well known, and that the measurements of rotor velocity and stator current are suciently accurate, the current-based estimator dynamics can be written as d r = dt r r M R R I + re J r + is , Lr Lr

(3.35)

r is the estimated rotor resistance. The dierence between the actual and where R estimated rotor resistance is dened as follows: r eRr = Rr R (3.36)

If there is a dierence between the actual and estimated rotor resistance, then there will also be a dierence between the actual and estimated rotor ux linkages. This error is dened as follows: er = r r The dynamics of the rotor ux linkage error can therefore be written as d d d er = r r dt dt dt = Rr Rr M I + re J r + is Lr Lr r r M R R I + re J r + is Lr Lr (3.38) (3.37)

Rr eRr eRr M I + re J er r + is Lr Lr Lr

We then choose our rotor resistance estimator to be the following: d Rr = K ir T er dt

(3.39)

41 M 1 r is and K > 0 is a constant control gain. where ir = Lr Lr Assuming the actual rotor resistance is changing very slowly d Rr 0 dt the dynamics of the rotor resistance error can be written as d eRr = K ir T er dt

(3.40)

(3.41)

To prove the stability and convergence of our rotor resistance estimator, we will use the Lyapunov stability criterion and Barbalats lemma. For a given K > 0, we choose the following Lyapunov function candidate:

T 2 V (er , eRr ) = K ||er ||2 + e2 Rr = Ker er + eRr

(3.42)

The rst derivative with respect to time can be computed as follows: d T e dt r d er dt d eRr dt

=K V

T er + er

+ 2eRr

(3.43)

The error dynamics can then be substituted into the expression: Rr eRr eRr M I + re J er r + is Lr Lr Lr
T

= K

er

T +Ker

Rr eRr eRr M I + re J er r + is + 2KeRr ir T er Lr Lr Lr 1 T M T + is Lr r Lr er eRr + ir T er eRr (3.44)

= 2K = 2K

Rr ||er ||2 + Lr

Rr ||er ||2 ir T er eRr + ir T er eRr Lr

42 Finally, we arrive at our result = 2K Rr ||er ||2 V Lr

(3.45)

Therefore, we may acknowledge at this point that our estimator is stable in the sense of Lyapunov, given this negative semidenite result. We desire however, to prove that our estimator is asymptotically stable and therefore the rotor ux linkage and rotor resistance errors will converge to zero. To do this, we will use Barbalats lemma as given in Applied Nonlinear Control by J.-J. E. Slotine [5]. As it applies to our problem, we will show that our Lyapunov function, V , which , which is dierentiable and has a nite limit as t , has a rst derivative, V 0 as t . To show that V is uniformly is uniformly continuous and so V , is bounded. The second derivative continuous, we will prove that its derivative, V of our Lyapunov function with respect to time can be computed as follows: = 2K Rr V Lr = 4K d T e dt r d er dt

T er + er

Rr Rr ||er ||2 eRr ir T er Lr Lr Rr Lr


2

= 4K

||er ||2 + 4K

Rr T eRr ir er Lr

(3.46)

we make the following observations: To determine the boundedness of V 0 V L since V Therefore er L and eRr L ir T L L and so V 0 er 0 as t . Therefore we may conclude that V

43

3.3.2

Steady State Analysis

While we have shown that the estimator is stable in the sense of Lyapunov and that the rotor ux linkage error will go to zero as time goes to innity, we have yet to show that the rotor resistance error will converge to zero as well. To show this, we will use a quasi-steady-state analysis. For this analysis to be valid, we much choose the adaptive estimator gain, K , such that the rotor resistance error dynamics are much slower than the electrical dynamics in the expression (3.41). It should also be noted that we have shown that the rotor ux linkage estimators are stable in previous sections. Therefore we can rewrite the rotor resistance error dynamics assuming the electrical dynamics are in steady state: d eRr = K Ir T r r = K Ir T r dt where M Ir is = Lr r se T r 1 + se T
2

(3.47)

r I + J I is se T s

(3.48)

r is = Lr Ir is + M Is is = M

1 se Tr is 2I 2 J Is 1 + (se Tr ) 1 + (se Tr )

(3.49)

We have used expressions written in the stator current reference frame, denoted by the superscript is , to obtain a result which describes the estimators dependency on the slip frequency. The use of this reference frame will not aect the nal results however, as the nal expression will depend on the magnitude of rotor ux linkage, which is the same for all reference frames. Substituting (3.48) and (3.49)

44 into (3.47), we obtain the following: 2 se Tr Tr Tr d M2 eRr = K dt Lr 1 + T 2 1 + se Tr se r where Is


2

Is

(3.50)

r Isd

r Isq

= 1 + se Tr

r M2

(3.51)

r Tr = Lr eRr and substituting (3.51) into (3.50), we obtain the Noting that T r Rr R desired expression: se Lr Lr Rr eRr = K Rr se Lr 1+ Rr eRr
2 2

d eRr dt

eRr

(3.52)

Inspection of this nonlinear result reveals an important performance limitation. When the slip frequency, se , goes to zero, the estimator will no longer work. To prove asymptotic stability of the rotor resistance error dynamics, we will again use the Lyapunov stability criterion. We choose the following Lyapunov function candidate: V = e2 Rr The rst derivative with respect to time can be computed as follows: se Lr Rr eRr = 2K Lr V Rr se Lr 1+ Rr eRr where r
2 2 2 eRr

(3.53)

(3.54)

< 0 for all values of eRr = 0 > 0. Inspection of (3.54) reveals that V

45 and se = 0, and so we may conclude that the rotor resistance estimators error dynamics are asymptotically stable under these conditions. Since a method for turning the adaptation o at low slip frequencies will be proposed in Chapter 5, the restriction on se is not a concern.

3.4

Simulink Simulation

Simulations were performed to verify the stability and performance of our controller using the parameters in Table 3.2. To keep simulation times reasonable, we neglected the inverter in our simulations and instead directly supplied the ideal sinusoidal command voltages to the induction machine model. Additionally, we used simulinks continuous time integrators and transfer functions to implement our controller. The simulink block diagrams have been provided in Figures 3.93.15. To better match the laboratory testing conditions the simulation results [Figure 3.8] provided are for a xed rotor velocity of 1200 RPM.

Machine Parameters [16]

Controller Parameters

Rs = 1.59 Rr = 1.86 Ls = 116.5 mH Lr = 116.7 mH M = 109.5 mH P =4 H = 0.8 kg m2 m2 B = 0.1 kg s Kp = 20 Ki = 200 Ka = 20 n = 2 = 0.15

Stator Resistance Rotor Resistance Stator Inductance Rotor Inductance Mutual Inductance Number of Poles Rotor Moment of Inertia Rotor Damping Constant Proportional Gain Integral Gain Adaptive Gain Int. Approx. Nat. Frequency Int. Approx. Damping Ratio

Table 3.2. Simulation parameters.

46

4 [N-m] 2 0 -2 0 2 Rr [Ohms] 1 0 0 0.4 0.2 0 0 200 vs [V] 0 -200 0 10 is [A] 0 -10 0


||r||cmd ||r|| True Rr Estimated Rr 3ph cmd load

10

10

||r|| [V-s]

10

10

4 5 6 Time [Seconds]

10

Figure 3.8. Simulation results.

47

1 Math Function 2 1 i_sd 3 w_re 2 i_sq


Rr i_sq lambda _rq lambda _rq_hat w_re I_sd lambda _rd lambda _rd_hat

Divide 1e-10 Constant

cos (theta _lambda _r)

3 sqrt u 2 Add 2 Math Function 1 Add 3 || lambda _r ||

Current Based Est

Math Function

Divide 1 2 sin (theta _lambda _r)

M1 M 1/Lr Add 7 1/Lr1 1/Lr M M


i_sd lambda _rd i_sq

i_r_hat

Add 6

1/Lr

Ka -Ka Dot Product 1 s Integrator 2 Rr_hat To Workspace

Add 4
e_lambda _r

4 v_a 5 v_b

v _a lambda _rq v _b

Add 5

Voltage Based Est

Figure 3.9. Simulink block diagram of resistance estimator and reference frame calculator.

1 i_sd Rs Rs Ls Ls

M M 1 Add lambda_rd

Lr/M s 3 v_a 4 v_b 2 i_sq Rs1 Rs Ls1 Ls


a d

Add1 Lr/M

s2+2*zeta*wns+wn^2 Add2 Int Approx

Mod. Clark Trans. M M1 2 Add3 lambda_rq

Lr/M s s2+2*zeta*wns+wn^2 Add5 Int Approx 1 Add4 Lr/M1

Figure 3.10. Simulink block diagram of voltage-based rotor ux linkage estimator.

48

1/Lr2 1/Lr Product5 1 s Product Product2 1 i_sd M/Lr M/Lr1 4 Rr_hat 3 w_re Product4 Add Integrator

1 lambda_rd_hat

1/Lr 1/Lr3

1 s Product1 Product3 2 i_sq M/Lr M/Lr Add1 Integrator1

2 lambda_rq_hat

Figure 3.11. Simulink block diagram of current-based rotor ux linkage estimator.

Inverse Park Trans 0.4 || lambda _r|| PI2 PI1


qr in out in out dr d

Inv Clark Trans


a d i_b b v _b i_c v _a i_a

1/((3*P/4)*(M/Lr)) Divide

in

out refd

PI3 1 t_em Tem_c To Workspace1 Torque Conv


refq

q c v _c w_r

Load_Torque

theta _r

Step IM_3ph_Model Park Trans


d dr q q b

Mod Clark Trans


d a

refd cos (theta _lambda _r) qr refq sin (theta _lambda _r)

i_sd i_sq w_re v _a || lambda _r || v _b

P/2 P/2

Rotor Flux -Linkage Est and Adaptive Rotor Resistance Est

Figure 3.12. Simulink block diagram of the direct eld-oriented controller.

49

Modified Clark Transform


1 a sqrt(3)/3 sqrt(3)/3 2 b 2*sqrt(3)/3 2*sqrt(3)/3 2 q 2 q 1 d 1 d

Inverse Clark Transform


1 a 1/2 1/2 2 b 3 c

sqrt(3)/2 sqrt(3)/2

Figure 3.13. Simulink block diagrams of the Modied (left ) and Inverse (right ) Clark transforms.

Park Transform
1 d Product 3 refd 2 qr 1 dr

Inverse Park Transform


1 dr Product 3 refd 1 d

Product 2

Product2

2 q

Product 3 4 refq

4 refq 2 qr

Product3

2 q

Product 1

Product1

Figure 3.14. Simulink block diagrams of the Park (left ) and Inverse Park (right ) transforms.

Kp Kp 1 in Ki Ki 1 s Integrator

1 out

Figure 3.15. Simulink block diagram of a PI regulator.

Chapter

Implementation and Experimental Verication


The implementation of our direct eld-oriented controller and adaptive rotor resistance estimator require the experimental determination of the induction machine parameters along with an accurate measurement of the dead time. This chapter will present the experimental methods used to determine these parameters. Specifically, an induction machine parameter estimation technique based on least-squares will be presented along with the resulting parameters. Additionally, the details regarding the implementation of our controller using dSpace will be presented along with experimental results.

4.1

Induction Machine Parameter Estimation

Traditionally, the electrical parameters of an induction machine have been computed using data obtained from the Locked-Rotor and No-Load tests [4]. There is, however, some debate over the accuracy of the parameters obtained from these tests. This is due in part to the nature of the tests in that the parameters

51 are computed based on two data points and are therefore more prone to experimental error. Additionally, the locked rotor test cannot be performed at nominal ux-levels due to current limitations, and so the results do not reect a normal operating condition. The technique we have used in our work uses multiple data points at a constant nominal ux level to determine values from which the specic machine parameters can be computed.

4.1.1

Parameter Estimation Technique

The parameter estimation technique we used is based on the stator current locus in the stator ux linkage reference frame [Figure 4.1]. When the direct and quadrature currents are plotted as a function of the slip frequency in this referM2 ence frame, they trace out a circle which has a radius r = 2 ||s || and center 2 Ls 1 1 Lr xo = + ||s ||. 2 Ls 2

s Isq

se

~ s I s
jj~ sjj Ls
se = 0

xo

se = 1

Lr ~ 2 jjsjj s Isd

Figure 4.1. Stator current locus as a function of slip frequency in the stator ux linkage reference frame.

52 Our technique ts experimental data to this circle using a recursive algorithm created by Dr. Heath Hofmann. The algorithm, which was implemented in Matlab, approaches the problem by starting with a range of guesses for the location of the center of current locus, x o . The algorithm then uses a loop to compute the leastsquares estimation of the square of the radius, r 2 (i), for each guess, x o (i). The norm-squared error, e(i), is then computed for each guess, from which the guess yielding the smallest error, x o (imin ), is chosen as the best estimate for the center of the locus, x o , along with the corresponding radius, r = sqrt( r2 (imin )). 1:
s s o,min ,x o,max ,N ): Algorithm Parameter Estimation(Isd ,Isq ,x

2: x o = linspace( xo,min , x o,max , N ); 3: for i = 1 : length( xo ) s 4: x = Isd x o (i); s 5: y = Isq ; s 6: d = ones(length(Isd ), 1); 2 2 7: b=x +y ; 8: r 2 (i) = inv(dT d)dT b; 9: e(i) = sum(( r2 x 2 y 2 )2 ); 10: end 11: [emin , imin ] = min(e) 12: x o = x o (imin ) 13: r = sqrt( r2 (imin )) Once the best estimates of the center location, x o , and radius, r , have been determined, the stator inductance can be computed: s = ||s || L x o r

(4.1)

Next, we make the standard assumption that the stator and rotor inductances are s = L r ) [4] and compute the leakage term: the same (i.e. L 2 = rL s ||s || L sx 2L o ||s ||

(4.2)

53 The mutual inductance can then be calculated:

= M

sL r L 2

(4.3)

To estimate an expected value for the rotor resistance, we will again use leastsquares: x = DT D
1

DT y

(4.4)

where D and y are m 1 column vectors whose j th elements are given by: ||s ||j s L sej

Dj =

s Isd j

(4.5)

yj =

2 ||s ||j ||s ||j M s sej + sej Isd sej j 2 Ls Ls

(4.6)

Finally, we can solve for the rotor resistance: 2 r = R x s L

(4.7)

4.1.2

Experimental Setup and Results

A motor-generator conguration [Figue 4.2] was used in our laboratory testing. The load dyne is used to regulate the rotor velocity of the machine being tested which creates a generating condition in dyne. That generated power ows back into the DC bus which is supplying the induction machine being tested. Thus, the power ows around in a loop, and only the losses due to imperfect eciency need to be replenished. The manufacturer ratings for the test and load dyne induction machines are provided in Table 4.1.

54

a b BUS b

Gate Drive Signals

Gate Drive Signals

User Inputs

User Inputs

Figure 4.2. Experimental setup.

Test Motor Manufacturer: US Motor No. Phases: 3 V/I: 230 V/4.7 A Frequency: 60 Hz Power: 1.5 Hp Nominal Speed: 1745 RPM Torque: 4.5 lb-ft

Load Dyne Motor Manufacturer: US Motor No. Phases: 3 V/I: 230 V/13.6 A Frequency: 60 Hz Power: 5 Hp Nominal Speed: 1775 RPM Torque: 15 lb-ft

Table 4.1. Manufacturer machine ratings.

Before we could record the data needed to estimate the 1.5 Hp test machines parameters, we rst needed to measure the stator winding resistance, using a digital multimeter, and the dead time. These parameters were needed to implement the voltage-based stator ux estimator, which was used to adjust the voltage lev-

55 els such that a constant stator ux linkage magnitude of 0.2 V-s was maintained. The procedure used to measure the dead time is outlined in the following section, and the Simulink block diagram used to generate the code for the dSpace microcontroller has been provided in Figure 4.3. It should be noted that the safety shutdown controllers have been cut out of the gure to save space.
f() Trigger 1 Va
a v_sa D_a Duty cycle a

Host Service #1 <ServiceName> Data Capture

2*pi*60 we

Electrical Frequency

vsd

2 Vb

v_sb D_b v_sc Duty cycle b

0.0 VllRMS

-KLine-line RMS to Line-neutral Peak

Voltage Magnitude

vsq

q c

3 Vc

theta_r

2-Phase Voltage Calculation

dq0-abc Conversion ab-dq Conversion1


d a

DeadTime Comp
Ia Vd_dt

V_bus

Dc

Duty cycle c

Rad_2_RPM
omega_r

Duty Cycle Generator 4 w_RPM Ref. Frame Trans.


d dr q Vbus PWM Stop

-K-

Encoder Interface
V_bus

Ib

DS1104SL_DSP_PWM3 <I_sd^Fs> 0.0001 6 z-0.9999 Avg. Filter1 ||Lambda_s|| i_sd 5 || lambda_s ||


i_sq refd cos(ls) v_sd refq sin(ls) v_sq

i_sa

Vq_dt DTC_ON 0 DTC_ON

i_sb

i_sc

<I_sq^Fs> 0.0001 7 z-0.9999 Avg. Filter

qr

V-Based Stator Flux Est. Measurements

Figure 4.3. Simulink block diagram for parameter estimation experiment.


0 STOP 0 Reset
Reset PWM Stop STOP

OR

PWM_stop

double Data Type Conversion1

STOP Shutdown

Once the dSpace microcontroller was programmed, we were able to take the necessary data. The electrical frequency was xed at 60 Hz and the dyne was
Overvoltage Shutdown
Reset a d PWM Stop Reset PWM Stop Vbus

Logical Operator

used to vary the rotor speed, which varied the slip frequency, from 1800 RPM
i_sd
I_sd Iq Peak Stator Current b q

Overcurrent Shutdown i_sq

down to 1720 RPM, in 10 RPM increments. As mentioned previously, the voltage


ab-dq Conversion

magnitude was adjusted to maintain a stator ux linkage magnitude of 0.2 V-s. The two-phase equivalent stator currents in the stator ux linkage reference frame were recorded for each rotor speed. This data was then entered into our recursive algorithm to obtain the test machine parameters in Table 4.2. A plot comparing the experimental data to the tted circle has also be provided along with the normalized error [Figure 4.4].

56
6 4 2
s Isq

Fitted Circle Experimental Data

0.07 0.06 0.05


e ||e||

0 2 4 6 0 5 10

0.04 0.03 0.02 0.01 5 6 7 8

s Isd

x o

Figure 4.4. Comparison of experimental data and tted stator current locus (left ) with normalized error (right ).

Rs = 1.2 Rr = 937 m Ls = 109.2 mH Lr = 109.2 mH M = 100.1 mH P =4

Stator Resistance (Measured) Rotor Resistance Stator Inductance Rotor Inductance Mutual Inductance Number of Poles

Table 4.2. Estimated test machine parameters.

4.2

Dead Time Measurement

As mentioned earlier, the IGBTs in the three-phase inverter are switched ON and OFF by what are referred to as gate drives. These gate drives have a built in dead time, td , to protect the IGBTs which they are connected to, and while maximum and minimum dead times are specied in the manufacturers data sheet, the range was too wide, and a more exact value was needed. To measure this dead time we connected a resistor and inductor in series with

57 a single phase of the inverter [Figure 4.5]. We then commanded a constant 20% duty cycle at a switching frequency of 10 kHz in dSpace and observed the output voltage, Vo , on an oscilloscope using a Tektronix High-Voltage probe [Figue 4.6]. Using the oscilloscope, we measured the pulse width of the voltage waveform and subtracted it from the ideal pulse width of 20 s to get a dead time of 3.88 s.

D high vge, high


BUS

74.7 vge, low D low


o

408.7 mH

Figure 4.5. Dead time measurement experimental setup.

Figure 4.6. Dead time measurement for a 20 percent duty cycle.

58

4.3

Controller Experimental Verication

To test the performance of our controller, we used the same motor-generator conguration [Figure 4.2] that was used to collect data for estimating the machine parameters. In our experiments, VBU S , was generated by rectifying a 208 VRM S threephase power supply using a diode bridge (not pictured). The variable-frequency AC voltage used to drive the induction machine was generated from the DC bus voltage using an IGBT bridge [Figure 4.7] and Conventional Pulse-Width Modulation (CPWM). Our control algorithm uses measurements of the stator currents (ia and ib ), provided by Hall-Eect sensors, and a measurement of the DC bus voltage, along with torque and rotor-ux commands, to generate the appropriate
+ ), which turn the IGBTs ON and OFF, using the CPWM . . . Dc duty cycles (Da

technique.

Da+

D b+

Dc+

BUS

Da-

Db-

Dc-

Figure 4.7. IGBT Bridge Conguration.

4.3.1

Implementation

As mentioned earlier, our control algorithm was implemented using dSpace, which is the name of the German company who produces a Rapid Control Prototyping

59 (RCP) hardware/software package designed to seamlessly interface with Matlab and Simulink. The hardware used in our experiments was the DS1104 R&D Controller Board which includes the signal processing hardware, such as D/A and A/D converters, needed to implement our control algorithm. A block diagram from the manufacturers brochure, which describes the controller boards architecture has been provided in Figure 4.8.

Figure 4.8. dSpace DS1104 block diagram [30].

A key feature of the dSpace package is the ability to program the controller board by converting a Simulink block diagram to C-code, which is then loaded onto the DS1104 controller board. This allowed us to copy our simulation block diagram [Figure 3.12] for the control algorithm into the triggered subsystem [Figure 4.9]. We then entered all of the machine and controller parameters into the various blocks, replaced the continuous-time integrators with discrete-time versions, and replaced the continuous-time ux linkage estimators with our trapezoidal versions.

60

f() Trigger 0.01 || lambda_r ||* 0 t_3ph*


||lambda_r||* t_3ph* i_a i_b V_bus theta_r Int_Reset w_r omega_r D_b v_c v_sc v_b v_sb v_sa v_a D_a

Host Service #1 <ServiceName> Data Capture

Duty cycle a

Duty cycle b

DFOC w/ Adaptive Rr Est.


V_bus D_c Duty cycle c

Encoder Interface 0 Int_Reset 0 STOP 0 Reset


Reset STOP PWM Stop

Duty Cycle Generator


PWM Stop

STOP Shutdown
Reset w_r PWM Stop

DS1104SL_DSP_PWM3

Overspeed Shutdown
Reset V_bus Vbus PWM Stop

OR

double Data Type Conv

Overvoltage Shutdown
i_sa a d

Logical Operator

i_sd

Reset I_sd PWM Stop Iq

i_sb

i_sq

Overcurrent Shutdown

Mod Clark Trans


i_sc

Current Measurements

Figure 4.9. Simulink block diagram of controller implementation for dSpace code generation.

The use of a triggered subsystem gives us the ability to execute our control algorithm in synchronism with the switching period of the power electronics. This was done by using the PWM interrupt signal from the DS1104 controller board to trigger our control algorithm, which was running in the subsystem, between switching transitions. This ensures that the dSpace control board does not sample during a high-voltage transient which can generate a signicant amount electromagnetic interference (EMI). The duty cycle commands sent out to the dSpace controller board were calcu-

61 lated using the following expression: vxn (t) 1 + , 2 VBU S

Dx (t) =

(4.8)

where vxn (t) is the desired line-to-neutral voltage computed by our control algorithm. While angular brackets indicate that the command voltage is an average value, sinusoidal voltages can be commanded as long as there is a sucient separation between the switching frequency, fs , and the frequency of the commanded fs AC voltage ,fxv . The general rule of thumb is that 20 [1]. The Simulink fxv block diagram for this computation has been provided in Figure 4.10.

1 v_sa Product1 2 v_sb Product2 3 v_sc Product3 4 V_bus 0.5 0.5

1 D_a

2 D_b

3 D_c

Figure 4.10. Simulink block diagram of duty cycle generator.

The rotor ux linkage estimators presented in Chapter 3 were implemented in Simulink using delay blocks. In the voltage-based estimator, the second-order continuous-time transfer function used to approximate an ideal integrator was replaced with a delay block implementation [Figures 4.11] which implements the trapezoidal form derived in Chapter 3. The coecients where calculated in a separate subsystem [Figures 4.12] to allow real-time access to the values of and n in the dSpace Control Desk.

62
Coef Calc
1+2zd+d^2 1+2zd-d^2 1-2zd-d^2 wn^2Ts

Product 1 Ts/2 Gain 2 Product 2 1 1 1 e_k (Ts^2)/2 Divide 1 Gain 1 Subtract 1 z Delay 1 Product Divide Subtract z Delay

1 lambda _s_k

Ts Gain

Figure 4.11. Simulink block diagram of second order integrator approximation for the voltage-based estimator.
3 1-2zd-d^2 Add2 zeta zeta u wn wn Gain Ts/2 d 2 Constant1 Product 2 Add1 1 1+2zd+d^2 1 2 Add Constant u Ts Gain1 4 wn^2Ts 2 1+2zd-d^2

Junk

Junk1

Figure 4.12. Simulink block diagram of coecient calculation for the second order integrator approximation.

Similarly, the trapezoidal derivation of the current-based estimator (3.27-3.29) was also implemented using a delay block implementation [Figure 4.13]. The equation was rewritten with a series of coecients [Figures 4.14-4.16] which multiply with the states of the estimator. This helped clean up the implementation, and allowed access to the rotor resistance value which was updated by the adaptive estimator.

63

Gain 2 1 1 I_sd 2 Product Product 1 Product 3 Product 6 Add Divide z Delay 1 lambda _rd

2 w_re
w_re[k] a[k] Rr w_re[k] b[k] Rr c[k] w_re[k] Rr d[k] w_re[k] Rr e[k]

a[k] calc 4 Rr

b[k] calc

c[k] calc

d[k] calc

e[k] calc

Product 2

Product 4 Product 7 1 Divide 1 z Delay 1 2 lambda _rq

3 i_sq

2 Gain 1 Product 5

Add1

Figure 4.13. Simulink block diagram of current-based estimator.


2 1 a[k] 1 w_re[k] 4*Ts Gain1 1 b[k]

1 w_re[k]

Ts Gain1 Ts/Lr Gain2

Math Function1 Add 4-u[1]^2 Fcn2

2 Rr

Figure 4.14. Simulink block diagrams of the current-based estimator coecients a[k] (left ) and b[k] (right ).

Ts/Lr Gain 1 Rr Ts*M/Lr

2+u[1] Fcn2 1 c[k]

1 w_re[k] 2 Rr Ts*Ts*M/Lr Gain1 Product1 1 d[k]

Product1

Gain1

Figure 4.15. Simulink block diagrams of the current-based estimator coecients c[k] (left ) and d[k] (right ).
1 w_re[k] Ts Gain1 Ts/Lr Gain u 2 1 e[k]

Math Function Add (2+u[1])^2 Fcn

2 Rr

Figure 4.16. Simulink block diagram of the current-based estimator coecient e[k].

64

4.3.2

Experimental Verication

The experimental results presented in this section were obtained using the machine parameters, hardware, and controller described in prior sections of this thesis. Figure 4.17 shows the experimental transient response of our adaptive rotor resistance estimator for initial values of 1.5 ohms (top ) and 0.5 ohms (bottom ). These responses were taken at a xed rotor velocity of 1200 RPM and with the following controller specications: Flux Regulator Gains:(Kp = 10, Ki = 100), Current Regulator Gains:(Kp = 20, Ki = 200), Ka = 20, = 0.2, n = 5 rad/sec. It should also be noted that sampling frequency of our experimental controller is always the same as the switching frequency, fs = 10 kHz.

Rr [Ohms]

1.5 1
Adaptation ON

Estimated Rr Expected R
r

0.5 0 0 2 0.5 1 1.5 2 2.5 3 3.5 4

Rr [Ohms]

1.5 1 0.5 0 0

Adaptation ON

0.5

1.5

2.5

3.5

Time [Seconds]
Figure 4.17. Adaptive estimator experimental transient responses.

65 To test the torque response capabilities of our controller, we adjusted the regulator gains to achieve a very fast transient response. The following controller specications were used: Flux Regulator Gains:(Kp = 25, Ki = 2000), Current Regulator Gains:(Kp = 30, Ki = 4000), Ka = 20, = 0.2, n = 5 rad/sec. The resulting plots for a 2 N-m torque step at a rotor velocity of 1200 RPM, have been provided in Figure 4.18. The top plot compares the calculated torque response (computed using equation 3.1) with the step command value (indicated by the subscript cmd ). The next plot compares the estimated rotor resistance with the expected value obtained from our parameter estimation experiments; inspection of which reveals that our resistance estimator remains stable under torque transients. We will see in the following chapter however, that the steady state value of this resistance estimation is signicantly aected by rotor velocity as well as the slip frequency. The next plot compares the estimated rotor ux linkage magnitude, as computed by the current-based ux estimator, with the commanded value. Inspection of this plot reveals that ux regulator is not aected by torque transients. Thus, the desired decoupling between ux and torque regulation was achieved. The remaining plots show the commanded stator voltages (the actual stator voltages were not measured) and the measured stator currents. The smooth sinusoidal nature of these currents is an indicator that our torque regulating direct eld-oriented controller is working well.

66

4 [N-m] 2 0 0 1.5 Rr [Ohms] 1 0.5 0 0 0.4 ||r|| [V-s] 0.2 0 0 200 vcmd [Volts] 0 -200 0 10 is [Amps] 0 -10 0 ||r||est ||r||cmd 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.005 0.01 0.015 0.02 0.025 0.03 Estimated Rr Expected Rr 0.035 0.04 3ph cmd 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04

0.005

0.01

0.015

0.02

0.025

0.03

0.035

0.04

0.005

0.01

0.015 0.02 0.025 Time [Seconds]

0.03

0.035

0.04

Figure 4.18. Experimental response to a torque step from 1 N-m to 3 N-m.

67 In order to verify that the adaptive rotor resistance estimator was working correctly, we compared true and estimated torque responses, which were computed using the following methods. The estimated torque response was computed using the equation from Chapter 3: 3P M (rd isq rq isd ) 4 Lr

3ph =

(4.9)

where rd and rq were provided by the current-based estimator. The true torque response was computed using the following equation: 3P (sd isq sq isd ) 4

3ph =

(4.10)

where sd and sq were provided by the voltage-based estimator. This computation was taken as truth because it does not depend on the rotor resistance. In practice, the estimation of the stator ux linkage using the voltage-based ux estimator is very reliable provided the electrical frequency is suciently high. The estimated torque response (referred to as the Rotor-Based Torque) and the true torque response (referred to as the Stator-Based Torque) were captured at a xed rotor velocity of 1200 RPM, and at a rotor ux linkage magnitude of 0.3 V-s. The adaptation was turned o during this test (i.e. Ka = 0), so that the rotor resistance used by the current-based estimator could be xed at values of 1.5 ohms and 0.5 ohms. Finally, the adaptation was turned on long enough to tune the rotor resistance value before being turned o for the nal comparison. The resulting plots for the mismatched and tuned resistances have been provided in Figure 4.19.

68

r = 1 .5 R 4 [Nm] 3 2 1 0 0 4 [Nm] 3 2 1 0 0 4 [Nm] 3 2 1 0 0 0.02 0.04 0.06 0.08 Time [Seconds] 0.1 RotorBased Torque StatorBased Torque Torque Command 0.12 0.14 0.02 0.04 0.06 0.08 r = 1 .04 (Adaptation Tuned) R 0.1 0.12 0.14 0.02 0.04 0.06 r = 0 .5 R 0.08 0.1 0.12 0.14

Figure 4.19. Comparison of torque steps for mismatched and tuned rotor resistances.

Inspection of the torque steps in Figure 4.19 reveals that our rotor resistance r , is estimator is meeting our objective. When the estimated rotor resistance, R xed at incorrect values of 1.5 ohms and 0.5 ohms, the mismatch results in a large steady-state error in the regulated torque (top two plots ). When our adaptive estimator is allowed to tune the rotor resistance, the regulated (rotor-based) torque matches the true (stator-based) torque very well (bottom plot ).

Chapter

Discussion and Future Work


This chapter will discuss the limitations of our controller and present experimental data pertaining to these limitations. Additionally, recommendations for further research will be made. The rst section will discuss the limiting factors as related to the rotor resistance estimators performance at various operating points, as well as the rotor ux linkage estimators performance over a range of frequencies. The nal section will propose acceptable operating points, based on the observed data, for tuning the rotor resistance estimation as well as considerations for improving the performance of the estimator.

5.1

Controller Performance and Limitations

Limiting factors have been discussed throughout this thesis. In this section we will present experimental data which veries these limitations. The limitations regarding the accuracy of our rotor resistance estimator will be presented rst, followed by data indicating the frequency limitations of our rotor ux linkage estimators.

70

5.1.1

Dependence on Operating Point

While testing our rotor resistance estimator, it became apparent that the resistance estimation had a dependency on the operating point of the induction machine as dened by the rotor speed and slip frequency [Figure 5.1]. Inspection of this plot reveals some interesting trends. The estimated resistance increases with rotor velocity; this could potentially be attributed to core losses, which tend to increase with frequency.

1.5

1 Est. Rr [Ohms]

r = 200 RPM 0.5 r = 600 RPM r = 1000 RPM r = 1400 RPM r = 1800 RPM Expected Rr 0 2 3 4 5 6 7 8 se [Rad/Sec] 9 10 11 12

Figure 5.1. Steady-state rotor resistance estimation at dierent rotor velocities as a function of slip frequency.

More interesting, is the decreasing trend of the resistance estimation at higher slip frequencies. Our expectation is that the rotor resistance would increase with slip frequency due to the skin-eect phenomenon in the rotor bars [1]. However, inspection of Figure 5.1 reveals that this is, in fact, not the case. It is expected that

71 some unmodeled eect, such as magnetic nonlinearity (saturation), is the cause of this trend. It is also possible that the trends in the estimator data could be caused by non-ideal eects steaming from the implementation of our rotor ux linkage estimators.

5.1.2

Flux Estimator Performance

The performance of our adaptive estimator depends, in part, on the performance of our rotor ux linkage estimators. Therefore, it was desirable to characterize their performance over a range of rotor velocities and slip frequencies. To do this, we ran the experimental setup with an open-loop controller, using the load dyne to x the rotor speed. The stator voltage commands were generated at a xed frequency and amplitude, which were adjusted to obtain the desired slip frequencies and rotor ux linkage magnitude (||r || = 0.3 V-s). The following dierences were then computed at each operating point: e||r || = ||r ||v ||r ||i T r,v r,i ||r,v || ||r,i || (5.1)

|e | = vi = cos1

(5.2)

where the subscript v indicates the value is obtained from the voltage-based estimator, and a subscript i indicates the value is obtained from the current-based estimator. It should be noted that the estimator parameters used in this test were the same as those used in the experimental verication. The resulting plots have been provided in Figure 5.2. Inspection of the rotor ux linkage magnitude dierence, e||r || , reveals that the voltage and current-based estimators are in relatively good agreement, particularly

72 above 1000 RPM. This trend was expected as the dead-time eect becomes less signicant at high speeds due to the increase in stator voltage magnitude, and so the voltage-based estimator becomes more accurate as the rotor velocity increases. The phase dierence, |e |, between the voltage and current-based estimators is particularly interesting. While it tends to reach a minimum around 600 RPM, it begins to increase again as the rotor velocity increases. While it is not clear at this time what mechanism is the cause of this trend, we suspect it could be the result our discrete-time implementation of the ux linkage estimators. Unmodeled time delays inherent to the DSP implementation could be contributing to the additional phase shifts.
0.06

= 2 Rad/Sec
se se se

= 4 Rad/Sec

e||r|| [V-s]

0.04

= 6 Rad/Sec se = 8 Rad/Sec = 10 Rad/Sec

0.02

se

se = 12 Rad/Sec

0 0 200 400 600 800 1000 1200 1400 1600 1800 2000

0.08 0.06 0.04 0.02 0 0

|e| [Radians]

200

400

600

800 1000 1200 Rotor Speed [RPM]

1400

1600

1800

2000

Figure 5.2. Open-loop characterization of agreement between the voltage and currentbased rotor ux linkage estimators.

73

5.2

Suggestions for Future Work

This thesis has presented a unique adaptive rotor resistance estimator in addition to rotor ux linkage estimators which yield improved performance over an impressive speed range. Both the resistance and ux linkage estimators were subjected to a thorough analysis and experimental evaluation of their performance at various operating points. The resulting direct eld-oriented controller, utilizing these estimators, has provided an acceptable level of performance over a wide range of these operating points. However, changes which remain to be investigated and implemented, will likely improve the controllers performance even further. One such improvement is the incorporation of magnetic saturation into the induction machine model.

ls

lr

re r

Figure 5.3. Induction machine model incorporating magnetic saturation eects.

While there are a variety of techniques which may be used to model magnetic saturation in an induction machine, the proposed method incorporates the saturation eect by making the mutual inductance a nonlinear function of the air-gap ux linkage magnitude [Figure 5.3]. This nonlinearity may be determined experimentally and then implemented using either a look-up table, or computed using a formula which has been tted to the experimental data.

74 Due to time constraints, a method for turning the rotor resistance adaptation on and o, was not tested. Based on the experimental transient responses [Figure 4.17], a technique which would simply switch the adaptive control gain between a value of 0 (o) and an acceptable value Ka > 0 (on), could be used without degrading the stability of the system. A better method, however, would be to use a weighting function to gradually adjust the adaptive gain. Both methods would a calculation of the electromagnetic torque, and a measurement of the rotor velocity to determine when the rotor resistance can be estimated accurately. It should be noted that torque has been suggested in place of slip frequency because it is more readily available. Based on the current performance, the acceptable operation points would likely be when the rotor speed is between 600 and 1000 RPM, and when the slip frequency is between 4 and 12 radians per second [Figure 5.1].

3ph

3ph

s a s

re s r

Figure 5.4. Block diagram of the proposed method for turning the resistance estimator ON and OFF.

75 Additional suggestions for future work include improving the method used to compensate for the dead-time eect, as well as exploring the feasibility of using ltered stator voltage measurements in the control algorithm. Also, improving the control design by attempting to model any additional time delays which have been overlooked or neglected, may yield improved performance as well. Finally, improvements to the induction machine model such as incorporating core loss in addition to magnetic saturation may prove worthwhile.

Bibliography

[1] H. F. Hofmann. AC Machines and Drives. The Pennsylvania State University, 2009. [2] B. K. Bose. Modern Power Electronics and AC Drives. Prentice Hall PTR, New Jersey, 2002. [3] P. C. Krause, O. Wasynczuk, S. D. Sudho. Analysis of Electric Machinery and Drive Systems. 2nd ed., John Wiley and Sons, New Jersey, 2002. [4] A. E. Fitzgerald, C. Kingsley, Jr., S. D. Umans. Electric Machinery. 6th ed., McGraw-Hill, New York, 2003. [5] J.-J. E. Slotine, W. Li. Applied Nonlinear Control. Prentice Hall, New Jersey, 1991. [6] H. K. Khalil. Nonlinear Systems. 3rd ed., Prentice Hall, New Jersey, 2002. [7] J. S. Bay. Fundamentals of Linear State Space Systems. WCB/McGraw-Hill, Boston, 1999. [8] C.-T. Chen. Linear System Theory and Design. Holt, Rinehart, and Winston, New York, 1984. [9] C. L. Phillips, H. T. Nagle. Digital Control System Analysis and Design. 3rd ed., Prentice Hall, New Jersey, 1995. [10] K. J. Astrom, B. Wittenmark. Adaptive Control. 2nd ed., Dover, New York, 2008. [11] P. Kokotovic, H. K. Khalil, J. OReilly. Singular Perturbation Methods in Control Analysis and Design. SIAM, Pennsylvania, 1999. [12] S. Venit, W. Bishop. Elementary Linear Algebra. 4th ed., Brooks/Cole, California, 1996.

77 [13] R. H. Park. Two-Reaction Theory of Synchronous Machines, Generalized Method of Analysis - Part 1, A.I.E.E. Transactions, vol. 48, 1929. [14] C. Kral, T. G. Habetler, R. G. Harley, F. Pirker, G. Pascoli, H. Oberguggenberger, and C.-J. M. Fenz. Rotor Temperature Estimation of Squirrel-Cage Induction Motors by Means of a Combined Scheme of Parameter Estimation and a Thermal Equivalent Model, IEEE Transactions on Industry Applications, vol.40, no.4, July/August 2004. [15] R. Krishnan, F. C. Doran. Study of Parameter Sensitivity in HighPerformance Inverter-Fed Induction Motor Drive Systems, IEEE Transactions on Industry Applications, vol.IA-23, no.4, July/August 1987. [16] H. F. Hofmann, S. R. Sanders. Speed-Sensorless Vector Torque Control of Induction Machines Using a Two-Time-Scale Approach, IEEE Transactions on Industry Applications, vol. 34, no. 1, pp. 169-177, January/February 1998. [17] P. L. Jansen, C. O. Thompson, and R. D. Lorenz. High Quality Torque Control at Zero and Very High Speed Operation, IEEE Industry Applications Magazine, July/August 1995. [18] G. Wang, H. F. Hofmann. Speed-Sensorless Torque Control of Induction Machine Based on Carrier Signal Injection and Smooth-Air-Gap Induction Machine Model, IEEE Transactions on Energy Conversion, vol. 21, no. 3, pp. 699-707, September 2006. [19] H. A. Toliyat, M. S. Arefeen, K. M. Rahman, and D. Figoli. Rotor Time Constant Updating Scheme for a Rotor Flux-Oriented Induction Motor Drive, IEEE Transactions on Power Electronics, vol. 14, no. 5, September 1999. [20] L.-C. Zai, C. L. DeMarco, T. A. Lipo. An Extended Kalman Filter Approach to Rotor Time Constant Measurement in PWM Induction Motor Drives, IEEE Transactions on Industry Applications, vol. 28, no. 1, January/February 1992. [21] D. P. Marcetic, S. N. Vukosavic. Speed-Sensorless AC Drives With the Rotor Time Constant Parameter Update, IEEE Transactions on Industrial Electronics, vol. 54, no. 5, October 2007. [22] Y. Koubaa and M. Boussak. Rotor Resistance Tuning for Indirect Stator Flux Oriented Induction Motor Drive Based on MRAS Scheme, European Transactions on Electrical Power, vol. 15, 2005. [23] R. Blasco-Gimenez, G. M. Asher, M. Sumner, K. J. Bradley. Dynamic Performance Limitations for MRAS Based Sensorless Induction Motor Drives.

78 Part 2: Online Parameter Tuning and Dynamic Performance Studies, IEE Proceedings-Electric Power Applications, vol. 143, no. 2, March 1996. [24] H.-J. Shieh, K.-K. Shyu, F.-J. Lin. Adaptive Estimation of Rotor Time Constant for Indirect Field-Oriented Induction Motor Drive, IEE ProceedingsElectric Power Applications, vol. 145, no. 2, March 1998. [25] Y. Zheng, H. A. Abdel Fattah and K. A. Loparo. Non-linear Adaptive Sliding Mode Observer-Controller Scheme for Induction Motors, International Journal of Adaptive Control and Signal Processing, vol. 14, 2000. [26] H. Kubota, K. Matsuse, T. Nakano. DSP-Based Speed Adaptive Flux Observer of Induction Motor, IEEE Transactions on Industry Applications, vol. 29, no. 2, March/April 1993. [27] H. Kubota, K. Matsuse. Speed Sensorless Field-Oriented Control of Induction Motor with Rotor Resistance Adaptation, IEEE Transactions on Industry Applications, vol. 30, no. 5, September/October 1994. [28] H. Kubota, D. Yoshihara, K. Matsuse. Rotor Resistance Adaptation for Sensorless Vector-Controlled Induction Machines, Electrical Engineering in Japan, vol. 125, no. 2, 1998. [29] H. Kubota, K. Matsuse, and T. Nakano. New Adaptive Flux Observer of Induction Motor for Wide Speed Range Motor Drives, Conf. Rec. IEEE IECON90, pp. 921-926. [30] DS1104 R&D Controller Board, Catalog 2009, dSPACE, www.dspace.com, pp. 332-337.

S-ar putea să vă placă și