Anda di halaman 1dari 17

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 44, NO.

5, MAY 1999

967

Global Adaptive Output Feedback Control of Induction Motors with Uncertain Rotor Resistance
Riccardo Marino, Sergei Peresada, and Patrizio Tomei
Abstract The authors design a global adaptive output feedback control for a fth-order model of induction motors, which guarantees asymptotic tracking of smooth speed references on the basis of speed and stator current measurements, for any initial condition and for any unknown constant value of torque load and rotor resistance. The proposed seventh-order nonlinear compensator generates estimates both for the unknown parameters (torque load and rotor resistance) and for the unmeasured state variables (rotor uxes); they converge to the corresponding true values under persistency of excitation which actually holds in typical operating conditions. In these cases the rotor ux modulus asymptotically tracks desired smooth reference signals which allows the motor to operate within saturation limits (so that modeling assumptions are met). As in eld-oriented control, the control algorithm generates references for the magnetizing ux component and for the torque component of stator current which lead to signicant simplications for current-fed motors. Simulations show that the proposed controller is suitable for high dynamic performance applications. This is conrmed by experiments since the control is robust with respect to modeling errors, sensors and actuators noise, control discretization, and simplication. Experimental comparisons with a classical indirect eld-oriented control exhibit a signicant improvement of power efciency when rotor resistance differs from its nominal value. Index Terms Adaptive nonlinear control, induction motor, online parameter estimation, output feedback.

I. INTRODUCTION

NDUCTION motors are more reliable and less expensive than those permanent magnet switched reluctance and d.c. motors which are currently employed for high dynamic performance applications; this is due to their simpler construction since they have no brushes, no commutator, no permanent magnet, and no windings in squirrel cage rotors. On the other hand, the control of induction motors is rather difcult. It is a highly coupled and nonlinear multivariable problem with two control inputs (stator voltages) and two output variables (rotor speed and ux modulus), required to track desired reference signals. Since ux sensors are not available, it is an output feedback problem in which the outputs to be controlled do not coincide with measured outputs (speed
Manuscript received May 16, 1997; revised June 30, 1998. Recommended by Associate Editor, J.-B. Pomet. This work was supported in part by MURST. R. Marino and P. Tomei are with the Dipartimento di Ingegneria Elettronica, Universit` a di Roma Tor Vergata, 00133 Roma, Italy. S. Peresada is with the Department of Electrical Engineering, Kiev Polytechnical Institute, Kiev 252056, Ukraine. Publisher Item Identier S 0018-9286(99)02825-1.

and stator currents). Moreover, there are uncertain critical parameters: in addition to load torque, which is typically unknown in all electric drives, in induction motors rotor resistance is also largely uncertain since it may vary up to 100% during operations, due to rotor heating. In comparison, the control of currently employed high dynamic performance motors is much simpler; this is in fact one of the reasons for their widespread use. However, the availability of low cost powerful digital signal processors (requiring only 40 ns to execute a simple instruction) and advances in power electronics make complex control algorithms now implementable even for medium- and small-size induction motors which could replace currently used motors, provided that high dynamic performance in positioning and speed tracking along with high-power efciency are achieved; this motivates current research efforts in induction motor control design. Assuming that all state variables (including rotor uxes) are available from measurements and all parameters (including rotor resistance) are known, the problem of controlling induction motors was solved by the so-called eld-oriented control [1], [2] and, more recently, by the inputoutput linearizing control [3] (see also [4] for a different feedback linearizing control and [5] for a variable structure approach). As claried in [6], both eld-oriented and inputoutput linearizing controls make use of nonlinear state-space change of coordinates and nonlinear state feedback transformations to achieve different objectives. While inputoutput linearization guarantees exact decoupling so that speed and ux modulus can be independently controlled with linear dynamics, eld-oriented control achieves the same properties only asymptotically, provided that the reference for the ux modulus is constant. This is a disadvantage especially in low power motors since at high speed, when the ux modulus reference has to be lowered (ux weakening) to operate the motor within the linear magnetic region and without voltage saturation or varied for torque maximization (see [7]), the speed tracking dynamics are perturbed. Advances in nonlinear adaptive control (see for instance [8]) allowed us to design in [6] an inputoutput state feedback linearizing control which is adaptive with respect to both load torque and rotor resistance. It is shown in [6] that load torque and rotor resistance estimation errors converge to zero in physical operating conditions. However, ux sensors are typically not available in induction motors since they reduce reliability and imply additional costs and technological difculties. Therefore, the design of state feedback controls is

00189286/99$10.00 1999 IEEE

968

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 44, NO. 5, MAY 1999

only a rst step toward the solution of the output feedback control of induction motors which points out two facts: high tracking performance is achieved while rotor resistance is estimated online; the resulting control is highly nonlinear but implementable with available microprocessors. The currently open problem, which is solved in this paper, is whether the same results may be obtained without ux measurements. Provided that rotor resistance is known, ux observers have been studied in [9][11] and exponentially convergent ux estimators are designed in [12] and [13]. They may be rather sensitive with respect to rotor resistance variations. Exponentially convergent ux observers have been tested in closed loop in experiments and simulations providing ux estimates both to eld-oriented [14], [2], [15] and inputoutput linearizing algorithms [16][20]. An independent line of research was focused on the design of rotor resistance estimators starting with [21] (see [22][28]), in which simplifying assumptions are made including linear approximations and steady state operations. More recently, an online nonlinear exponentially convergent rotor resistance estimator and an adaptive ux observer have been obtained in [29] which shows that both rotor uxes and rotor resistance may be estimated on line from speed and stator current measurements. An output feedback algorithm consisting of a ux observer which is adaptive with respect to rotor resistance together with a state feedback controller which is adaptive with respect to torque load is given in [30]; however, the closed-loop behavior is only illustrated by simulations since its stability is not proved in [30]. In fact, as far as nonlinear systems are concerned, replacing state measurements by exponentially convergent state estimates in global tracking controls does not yield, in general, global output feedback tracking controls. This property, which holds for linear systems, has to be independently established for nonlinear ones. Therefore the output feedback control problem for induction motors was addressed as such in [31][36], assuming known rotor resistance. In [31] speed and ux tracking are achieved by output feedback provided that the initial conditions lie in a predetermined region. In the series of papers [33][36], torque and ux tracking are globally obtained by output feedback using a passivity approach. As far as current-fed induction motors are concerned, an output feedback controller is presented in [37] which is free of singularities and experimentally gives high performance both in speed and ux tracking: however, it is shown in [37] that errors in rotor resistance reduce power efciency since higher currents are needed. Hence, it is proposed in [37] to update the rotor resistance value used by the controller by a rotor resistance estimator; nevertheless, since no stability proof has been obtained in [37] for the closed-loop system with continuously updated rotor resistance, the design of a global output feedback control which is adaptive with respect to rotor resistance, is still an open problem. Very recently, an adaptive output feedback position tracking controller is proposed in [38]; position tracking is achieved despite unknown inertia, load, and rotor resistance. The controller exhibits a singularity when the magnitude of the estimated rotor ux is zero and does not provide converging estimates for rotor resistance and rotor ux so that the rotor ux tracking

objective (which is crucial to improve power efciency) is not fullled. The contribution of this paper is to design a global output feedback control for induction motors which is adaptive with respect to both load torque and rotor resistance and guarantees asymptotic tracking of smooth speed references for any initial condition of the closed-loop motor. The resulting control is a seventh-order dynamic compensator which provides voltage inputs on the basis of rotor speed and stator currents measurements. The dynamic control algorithm generates estimates for load torque, rotor resistance, and rotor uxes that converge to the corresponding true values under persistency of excitation, which actually holds in physical operating conditions. The control also generates as internal signals stator current estimates and an estimate for the rotor ux angle. When persistency of excitation condition is satised the rotor ux modulus asymptotically tracks desired smooth reference signals, so that the motor operates within saturation limits at higher speed and the modeling assumptions (linear magnetic circuits) are met. As in classical eld-oriented control, reference signals both for the direct and the quadrature components of stator currents (in a frame attached to the estimated rotating ux vector) are generated which are responsible for ux modulus and speed tracking, respectively. Direct and quadrature current errors are forced to zero so that simplied controls may be used in current-fed machines relying on high gain current loops. Simulations and experiments show converging estimates both for rotor ux and for unknown parameters within 2 s and very precise tracking of speed and ux modulus reference signals, which are typically required in high-performance applications in spite of sharp load torque variations and rotor resistance uncertainties. Experiments conrm that since the proposed control algorithm achieves rotor resistance estimation and ux modulus tracking, an improved power efciency is obtained in comparison with a classical control scheme such as the indirect eld-oriented control. II. ADAPTIVE OUTPUT FEEDBACK CONTROL DESIGN A. Problem Statement Assuming linear magnetic circuits, i.e., no magnetic saturation, the dynamics of a balanced induction motor in a xed reference frame attached to the stator are given by the fthorder model (see for instance [6] for its derivation and [14] and [39] for modeling assumptions)

(1) in which the state variables are rotor speed , rotor uxes and stator currents ; the control inputs are

MARINO et al.: GLOBAL ADAPTIVE OUTPUT FEEDBACK

969

stator voltages ; the outputs to be controlled are rotor and rotor ux modulus ; the measured speed while the state variables are variables are , moment of not measured; the parameters are torque load and inertia , rotor and stator winding resistances mutual inductance ; the number of inductances pole pairs is equal to one. To simplify notations we use in (1) the parameters . The parameters and (or ) are typically uncertain. may in fact vary up to 100% of its nominal value due to rotor heating while depends on applications. whose dynamics will be later We introduce an angle dened

to design a dynamic output feedback compensator [recall that the measured outputs are ]

(5) so that, for any unknown by choosing and for any initial condition we have and

(6) and

(7) as a function of measured variables ( is an arbitrary initial condition). We also introduce the variables which imply that (8) asymptotNote that (7) implies that the ux vector , i.e., eld orientation is achieved. ically rotates at speed frame rotating at speed tends to In other words, the have the -axis coincident with the rotating ux vector as goes to innity. B. Nonadaptive Design Before proceeding in the design of the global adaptive control let us rst see how to design a global (nonadaptive) dynamic speed feedback controller when we restrict ourselves as the control to the rst three equations in (3) viewing inputs (current-fed motor model) and when we assume that and are known. As a matter of fact, in current-fed both are the control inputs to be designed. induction motors Our goal (6), (7) is achieved by choosing

(2) which represent the components of rotor ux, stator current and stator voltage vectors, respectively, with respect to a timeframe rotating at speed and identied by varying . In the new state coordinates the angle and new control coordinates the motor dynamics (1) become (see for instance [39] for induction machine modeling in arbitrarily rotating reference frames)

(9) (3) and the smooth bounded referLet us denote by ence signals for the output variables to be controlled, which and the rotor ux modulus are the speed , respectively, and by (10) Computing the time derivative of the candidate Lyapunov function ( is a positive design parameter) (11) and suitable functions . In fact, substitutwith ing (9) in the rst three equations in (3), we obtain

(4) ux components. the tracking errors of the speed and the Following the strategy of eld-oriented control [1], our goal is

970

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 44, NO. 5, MAY 1999

we obtain

for

and

, and the speed of the rotating

frame (18)

(12) Choosing (13) we have

is a positive design parameter and in which are additional terms yet to be designed. As we shall see, will be designed using a projection the adaptation for for all so that (17) is algorithm guaranteeing well dened. Introducing the current tracking errors (19) from (3), (17)(19) we obtain

which implies that is a globally exponentially stable equilibrium point for the closed-loop system (9), (10), (13) and (6), (7) hold for any initial condition and for any . In summary, the rstorder dynamic compensator

(20) (14) with given by The speed and ux tracking errors dynamics are

(21) (15) is a global output feedback tracking control for the rst three equations in (1) which represent the model of a current-fed induction motor. Substituting (13) in (10) we obtain and denote the parameter where in estimation errors. Note that if we set (21), we reobtain (10). It will be the goal of the adaptation to drive and to law for the parameter estimates and will zero while the objective of the control inputs and to zero. be to drive C. Adaptive Design (16) which shows how the closed-loop error dynamics are coupled in the ideal case of third-order reduced model with known and small gives well tuned parameters: choosing large and . Comparing the new responses in the tracking of global rst-order controller (14), (15) with the second-order one given in [37], we note that the main difference is that the latter makes use of a second-order ux observer. As an intermediate step in the design of the adaptive control, . Since we dene reference signals for the currents and are unknown constants, we modify the expressions and by their estimates and , (9) by replacing respectively. We dene the reference current signals the stator current estimates and by the corresponding estimation errors, we introduce the stator current observer Denoting by

(22) in a positive design parameter and are additional terms yet to be chosen. The stator current estimation error dynamics may be written as which is

(17)

(23) and will be helpful in the design of the adaptive law for .

MARINO et al.: GLOBAL ADAPTIVE OUTPUT FEEDBACK

971

We replace the unknown variables (recall that are not measured) by the new error variables (24) so that the error equations (21) and (22) are expressed in new coordinates as

We compute from (3), (17), (25), and (26) the dynamics of the stator current tracking errors

(25) instead The advantage of using the unknown variables relies on the fact that their dynamics no longer of depend on [compare (25) with (21)]. We now dene some and of the yet undetermined terms in (17), (18), (22) as ( in (22) are still to be chosen)

(28)

(26) positive design parameter and estimates of with the unknown error variables dened by (24). Note that the dynamics of are yet to be dened. If converge to , since they allow us to recover the rotor ux vector are known variables and is a known parameter. Hence, the may be viewed as rotor ux estimates. new variables Substituting (26) in (25), we obtain

with

(27)

972

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 44, NO. 5, MAY 1999

(assumed to be known). Alternatively, we may choose and with the maximum value of (assumed to be known). From (27) and (30), the time derivative of (30) results in

(31)

We now consider the function

(32)

From (31), (27), and (28), its time derivative is

(29)

, the dynamics of the estimaIn order to determine , and the feedback control inputs tion variables , we consider the function

(33)

(30) in which with chosen so that are positive parameters and are estimation errors. We dene and to guarantee that is with the minimum value of

, the feedback controls , and the The terms are now chosen in order to force dynamics of to be negative semidenite. Choosing the yet undetermined terms in (22)

(34)

MARINO et al.: GLOBAL ADAPTIVE OUTPUT FEEDBACK

973

we obtain

every

we modify the dynamics (37) according to

(39) is the smooth projection algorithm given in where [40] and dened in our case by if if

and otherwise

(35) We nally dene

with the minimum in which such that . The (known) value of and in (39) is chosen so that . The initial condition projection algorithm has the following properties: ; 1) is Lipschitz continuous; 2) ; 3) . 4) Property 4) implies that substituting (39) instead of (37) in (35), we obtain instead of (38) the inequality (40) From (32) and (40), it follows that are bounded for every . Their bounds depend on the initial errors. Therefore, and, consequently, according to (36), are bounded for every . Hence, are bounded and therefore are uniformly continuous. On the other hand, integrating (40) we have

(36)

(37) so that (35) becomes (38) Equations (36), (37), (29), (22), (26), (34), (17), (18), and (2) dene a seventh-order dynamic feedback compensator, whose , which generates state variables are on the basis of the measurements the control signals , the reference signals and their time deriva. In order to guarantee that for tives which implies by Barbalats lemma (see for instance [41] and [8]) that

This shows that asymptotic speed tracking is achieved from in (39). Moreover, any initial condition provided that and current tracking errors current estimation errors asymptotically tend to zero.

974

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 44, NO. 5, MAY 1999

D. Parameter Convergence We now analyze under which conditions also tend asymptotically to zero. The error equations are

(41) Whenever projection does not occur, i.e., may be rewritten as , (41)

(42) with the expressions shown at the bottom of the page. Making the change of coordinates and

with

, from (42) we obtain

Since the structure of

and

is such that

setting

we nally have

MARINO et al.: GLOBAL ADAPTIVE OUTPUT FEEDBACK

975

The radially unbounded function (32) may be written as , whose time derivative is . Since is skew-symmetric and recalland , we obtain , with ing the structure of . From now on, using the same arguments adopted in the proof of the persistency of excitation Lemma B.2.3 in [8, p. 367] we can establish that if persistency of excitation conditions are satised, i.e., there exist two positive constants and such that

III. CONTROL IMPLEMENTATION AND PERFORMANCE A. Control Implementation Let us rst summarize the seventh-order dynamic feedback control algorithm designed in Section II

(43)

of system (42) is unithen the equilibrium point formly asymptotically stable and all trajectories tend asymptotically to zero. If projection occurs, it is easy to see that the arguments in the proof of [8, Lemma B.2.3] are still valid and, therefore, we can conclude that if (43) holds then the equilibrium of system (41) point is uniformly asymptotically stable and all trajectories tend . In summary, asymptotically to zero provided that if (43) is satised then, in addition to asymptotic speed tracking: and tend to zero, from (24) 1) since both it follows that rotor ux traking is achieved and, in addition, the rotor ux vector is asymptotically oriented frame, i.e., with respect to the

2) since torque load

tend to zero, both rotor resistance are asymptotically estimated, i.e.,

and

(45) with 3) since tend to zero, the rotor ux vector is asymptotically estimated: dening, according to (24) given in (26), (34) and , given in (29). The dynamic compensator (45) contains eight control pawhose role may be evalrameters uated by examining both the closed-loop error equations (41) and the corresponding function (32) with time derivative (40). determine the rate of decay of The parameters in (40) and directly affect [see (41)] the dynamics of and speed tracking error , current estimation errors , respectively. The parameter current tracking errors determines the inuence of speed tracking errors on the other errors and is typically chosen much smaller than one. and may be separately tuned using The parameters has the desired transients. (16), (9), and (13) so that and are the adaptation gains for The parameters and , respectively. The smaller they are chosen, the

(44) we have in fact

976

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 44, NO. 5, MAY 1999

slower the adaptations for and result. The parameter is a weighting factor in the function (32); the choice of depends on the interval of variation of should be the the uncertain parameter . The parameter last parameter to be tuned and should be chosen sufciently large so that the current error dynamics are much faster than speed error dynamics while voltages are within saturation limits. Of course, different tuning strategies should be followed depending on sensor noise features, discretization strategies, and other implementation issues. For induction motors which allow for high gain current loops the control may be greatly simplied as follows:

(46) obtained by setting with (22), (26), (34), (36), and (39) in (17), (18),

(47) B. Control Performance We tested the proposed controller (45) both by simulations and by experiments [using its simplied version (46), (47)] for

a three-phase single pole pair 0.6-kW induction motor (OEMER 7-80/C), whose parameters are listed in the Appendix (see [29] for experimental and computed static speed-torque characteristics). Since ux and torque measurements are not available during the experiments, we rst tested by simulations how a typical ux modulus reference (including ux weakening) is tracked and the role of torque in persistency of excitation condition (43) when the controller (45) is used. We then illustrate by experiments the robustness of the simplied controller (46), (47) with respect to sensors noise, inaccuracies on mechanical and electrical parameters, control signals distortions generated by the power inverter, control discretization and truncation errors, inverter unmodeled dynamics, and unmodeled saturation effects of the magnetic circuits. Finally, we compare performances achieved by the proposed control with those given by a classical indirect eld-oriented control. The proposed control algorithm (45) has been tested rst by simulation with the control parameters (all values are in SI units): . Recall that may be chosen is negative. All initial conditions negative provided that of the motor and of the controller are set to zero excepting Wb and s , which is 50% greater s . The references than the true parameter value for speed and ux modulus along with the applied torque are reported in Fig. 1. The ux reference starts from 0.01 Wb at and grows up to the rated constant value 1.16 Wb; s. The speed reference is eld weakening starts at s and grows up to the constant value 100 zero until s the speed is required to go up to the value rad/s; at 200 rad/s, while the reference for the ux is reduced to 0.5 Wb. A constant load torque (5.8 Nm, the rated value) which s and reduced is unknown to the controller is applied at s. Fig. 2 shows the time histories of speed, to 0.5 Nm at and , the estimate ux modulus, load torque estimate . The speed tracks tightly the reference even of s and s), though load torque sharply changes (at since the load torque estimate quickly recovers the applied unknown value. Also the estimate of converges within 1 s to the true value. Note that the higher the torque is (see Fig. 3), the larger the convergence rate for is. Persistency of excitation condition (43) has been checked to hold. The ux tracks its reference: there is, however, a coupling with speed s and at s when speed is perturbed tracking at by an unknown load torque. Fig. 3 shows the time histories of torque, direct component of current estimation error, phasecurrent, and phase- voltage. Currents and voltages are within saturation limits and reduced overshoots are noticed in torque s and when unknown load torques are applied at s. The current estimation errors tend rapidly to zero after perturbations due to speed tracking errors. The simplied control algorithm (46), (47) was then tested experimentally with the control parameters values: ; the gains of the PI controllers (46) are chosen so that a unit step reference is tracked with a settling time of about 2.5 ms; all initial conditions of the controller . The following typical are set equal to zero excepting

MARINO et al.: GLOBAL ADAPTIVE OUTPUT FEEDBACK

977

Fig. 1. Reference signals and load torque in simulations.

Fig. 2. Speed, ux modulus, and parameter estimates in simulations.

operating conditions were experimentally tested: the unloaded motor is required to reach the rated speed 100 rad/s with acceleration 1000 rad/s in 140 ms starting from 0.5 s; during s, the motor ux modulus the initial time interval Wb to its rated value is driven from the initial value 10 1.16 Wb, with ux speed 3.87 Wb/s; both speed and ux reference signals (given in Fig. 4) are twice differentiable with

bounded second-order derivatives (the bounds are rad/s and 38.7 Wb/s , respectively); after start-up a constant load torque, equal to the rated value (5.8 Nm) is applied. Speed measurements are provided by an optical incremental encoder with 2000 lines per revolution, while current measurements are ltered by low pass lters with cut off frequency equal to 2.6 kHz and then converted by 12 bit A/D converters

978

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 44, NO. 5, MAY 1999

id ; ia ; and Fig. 3. Torque, ~

a in simulations.

Fig. 4. Reference signals in experiments.

with conversion time of 50 s. A 32-bit DSP (AT&T 32C) performs data acquisition, implements the control law using an improved Euler integration algorithm with a sampling time equal to 0.5 ms, and generates reference voltages for the power inverter with symmetrical PWM and switching frequency of 15 kHz. The DSP is hosted by a PC which programs the DSP, generates smooth speed and ux modulus reference signals, generates torque commands for a current controlled d.c. motor (which is connected to the induction motor), and stores and displays experimental data. We performed two experiments: underestimates the correct value , i.e., in the rst one ; while in the second one is overestimated, i.e., . The closed-loop performance in the two cases is documented in Figs. 5 and 6, respectively, in which speed , current estimation errors , estimated ux error modulus rents , voltage with given by (44), cur, and the normalized estimate

are given. In both cases speed errors are compatible with a high-performance drive; estimated ux modulus converges to converges to 1 within 2 s; the the reference value and estimation of ux modulus and of depends on the torque level (which may be evaluated from ). Finally, we performed for comparison the same two experiments by using the indirect eld-oriented control (FOC), which is implemented by (46) with

(48) where is used in the last equation instead of the usual

MARINO et al.: GLOBAL ADAPTIVE OUTPUT FEEDBACK

979

Fig. 5(a).

2 (t) + ^2 (t))1=2 ; i (t) and i (t) with initial underestimated rotor resistance. Experimental ! ~ (t); ( ^d q d q

Fig. 5(b). Experimental

uq ; ^ (t)= ; ~ iq (t)

id (t) with initial underestimated rotor resistance. and ~


observer (which is converging outside the magnetic saturation region)

estimated ux modulus. Comparing (48) with (47) we note that (48) may be viewed as a simplication of (47) and therefore with the same tuning can be used ( and ) which guarantees satisfactory speed in (48). The ux is estimated by the tracking when

980

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 44, NO. 5, MAY 1999

Fig. 6(a).

2 (t) + ^2 (t))1=2 ; i (t) and i (t) with initial overestimated rotor resistance. Experimental ! ~ (t); ( ^d q d q

Fig. 6(b). Experimental

uq ; ^ (t)= ; ~ iq (t)

id (t) with initial overestimated rotor resistance. and ~


(when compared with the corresponding in Figs. 5 and 6) are required to produce the rated torque: this is due to magnetic and to low ux modulus when saturation when . Experiments show that the controller proposed in this paper gives improved transient performance and a

which makes use of the true value . The performance achieved by FOC in the two cases are reported in Figs. 7 and 8; while the speed error is still satisfactory, the ux modulus and goes above the reference rated value when . In both cases higher currents below when

MARINO et al.: GLOBAL ADAPTIVE OUTPUT FEEDBACK

981

2 (t) + ^2 (t))1=2 ; i (t) and i (t) with FOC and underestimated rotor resistance. Fig. 7. Experimental ! ~ (t); ( ^d q d q

2 (t) + ^2 (t))1=2 ; i (t) and i (t) with FOC and overestimated rotor resistance. Fig. 8. Experimental ! ~ (t); ( ^d q d q

power efciency which is bigger than the one obtained by the indirect eld-oriented control (46), (48) with an inaccurate rotor resistance estimate. IV. CONCLUSION We have designed for the fth-order model (1) of an induction motor with constant load torque a seventh-order

nonlinear adaptive control (45) which, on the basis of rotor speed and stator currents measurements, guarantees asymptotic tracking of smooth speed references for any initial condition and for any unknown constant value of load torque and rotor resistance. Under persistency of excitation condition (43), we have shown that smooth rotor ux modulus reference signals

982

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 44, NO. 5, MAY 1999

are asymptotically tracked while the rotor ux vector tends to frame, achieving the coincide with the -axis of the so-called eld orientation; both unknown parameters (rotor resistance and load torque) and the rotor ux vector are asymptotically estimated. The control algorithm structure leads to a straightforward simplication for current-fed motors given by (46), (47). Simulations of typical operating conditions show accurate tracking of speed and ux modulus smooth reference signals usually required in high-performance applications in spite of load torque perturbations and rotor resistance variations, converging estimates of uncertain parameters and unmeasured state variables within 2 s. The simplied version (46), (47) of the controller was experimentally tested and compared with the classical indirect eld-oriented control (46), (48); simulation results were conrmed showing robustness with respect to sensor noise, unmodeled dynamics and saturation effects, inverter distortions, parameter inaccuracies, control discretization, and simplication. Moreover, improved power efciency and transient performance are documented in comparison with the classical indirect eld-oriented control algorithm when rotor resistance differs from its nominal value. APPENDIX MOTOR PARAMETERS Rated power Rated speed Rated torque Rated frequency Excitaton current Rated current Stator resistance Rotor resistance Mutual inductance Rotor inductance Stator inductance Motor-load inertia 600 W. 1000 rev/min. 5.8 Nm. 16.7 Hz. 2 A. 4 A.

[10] [11] [12] [13] [14] [15] [16] [17] [18] [19]

[20] [21] [22] [23] [24] [25] [26] [27]

H. H. H. kg m . REFERENCES

[1] F. Blaschke, The principle of eld orientation applied to the new transvector closed-loop control system for rotating eld machines, Siemens-Rev., vol. 39, pp. 217220, 1972. [2] W. Leonhard, Microcomputer control of high dynamic performance ac-drivesA survey, Automatica, vol. 22, pp. 119, 1986. [3] Z. Krzeminski, Nonlinear control of induction motors, in 10th IFAC World Congr., Munich, Germany. McGraw-Hill, 1987, pp. 349354. [4] A. D. Luca and G. Ulivi, Design of exact nonlinear controller for induction motors, IEEE Trans. Automat. Contr., vol. 34, pp. 13041307, 1989. [5] A. Sabanovic and D. Izosimov, Application of sliding modes to induction motor control, IEEE Trans. Ind. Applicat., vol. 17, pp. 4149, 1981. [6] R. Marino, S. Peresada, and P. Valigi, Adaptive input-output linearizing control of induction motors, IEEE Trans. Automat. Contr., vol. 38, pp. 208221, 1993. [7] M. Bodson, J. Chiasson, and R. Novotnak, A systematic approach to selecting ux references for torque maximization in induction motors, IEEE Trans. Contr. Syst. Technol., vol. 13, pp. 388397, 1995. [8] R. Marino and P. Tomei, Nonlinear Control DesignGeometric, Adaptive and Robust. London, U.K.: Prentice Hall, 1995. [9] Y. Dote, Existence of limit cycle and stabilization of induction motor via new nonlinear state observer, IEEE Trans. Automat. Contr., vol. 24, pp. 421428, 1979.

[28] [29] [30] [31] [32] [33] [34] [35] [36]

, Stabilization of controlled current induction motor drive system via new nonlinear state observers, IEEE Trans. Indust. Electron., vol. 27, pp. 7781, 1980. Y. Hori, V. Cotter, and Y. Kaya, A novel induction machine ux observer and its application to a high performance AC drive system, in 10th IFAC World Congr., Munich, Germany, 1987, pp. 363368. A. Bellini, G. Figalli, and G. Ulivi, Analysis and design of a microcomputer-based observer for an induction machine, Automatica, vol. 24, pp. 549555, 1988. G. C. Verghese and S. R. Sanders, Observers for ux estimation in induction machines, IEEE Trans. Indust. Electron., vol. 35, pp. 8594, 1988. W. Leonhard, Control of Electrical Drives. Berlin, Germany: SpringerVerlag, 1985. Y. Hori and T. Umeno, Flux observer based eld orientation FOFO controller for high performance torque control, in Proc. IPEC, Tokyo, Japan, 1990, pp. 12191226. D. I. Kim, I. J. Ha, and M. S. Ko, Control of induction motors via feedback linearization with input-output decoupling, Int. J. Contr., vol. 51, pp. 863883, 1990. , Control of induction motors for both high dynamic performance and high power efciency, Proc. Inst. Elec. Eng., 1992, vol. 139, pt., D, pp. 363370. R. Marino and P. Valigi, Nonlinear control of induction motors: A simulation study, in European Control Conf., Grenoble, France, 1991, pp. 10571062. T. von Raumer, J. M. Dion, L. Dugard, and J. L. Thomas, Applied nonlinear control of an induction motor using digital signal processing, IEEE Trans. Contr. Syst. Technol., vol. 2, pp. 327335, 1994. M. Bodson, J. Chiasson, and R. Novotnak, High performance nonlinear induction motor control via input-output linearization, IEEE Contr. Syst. Mag., vol. 14, pp. 2533, 1994. L. J. Garces, Parameter adaptation for the speed-controlled static AC drive with a squirrel-cage induction motor, IEEE Trans. Ind. Applicat., vol. 16, pp. 173178, 1980. T. Orlawska-Kowalska, Application of extended Luenberger observer for ux and rotor time-constant estimation in induction motor drives, Proc. Inst. Elec. Eng., vol. 136, pt. D, pp. 324330, 1989. T. Matsuo and T. A. Lipo, A rotor parameter identication scheme for vector controlled induction motor drives, IEEE Trans. Ind. Applicat., vol. 21, pp. 624632, 1985. L. C. Zai and T. A. Lipo, An extended Kalman lter approach in rotor time constant measurement in pwm induction motor drives, in IEEE Industry Appl. Soc. Ann. Mtg., Atlanta, GA, 1987, pp. 177183. D. Atkinson, P. Acarnley, and J. Finch, Observers for induction motor state and parameter estimation, IEEE Trans. Ind. Applicat., vol. 27, pp. 11191127, 1991. C. C. Chang and H. Wang, An efcient method for rotor resistance identication for high-performance induction motor vector control, IEEE Trans. Indust. Electron., vol. 37, pp. 477482, 1990. S. K. Sul, A novel technique of rotor resistance estimation considering variation of mutual inductance, IEEE Trans. Ind. Applicat., vol. 25, pp. 578587, 1989. J. Holtz and T. Thimm, Identication of the machine parameters in a vector-controlled induction motor drive, IEEE Trans. Ind. Applicat., vol. 27, pp. 11111118, 1991. R. Marino, S. Peresada, and P. Tomei, Exponentially convergent rotor resistance estimation for induction motors, IEEE Trans. Ind. Electron., vol. 42, pp. 508515, 1995. , Adaptive observer-based control of induction motors with unknown rotor resistance, Int. J. Adaptive Contr. Signal Proc., vol. 10, no. 4/5, pp. 345363, 1996. I. Kanellakopoulos, P. T. Krein, and F. Disilvestro, Nonlinear uxobserver-based control of induction motors, in American Control Conf., Chicago, IL, 1992, pp. 17001704. R. Marino, S. Peresada, and P. Tomei, Adaptive output feedback control of current-fed induction motors, in Proc. 12h IFAC World Congr., Sydney, Australia, 1993, pp. 451454. R. Ortega and G. Espinosa, Torque regulation of induction motors, Automatica, vol. 29, pp. 621633, 1993. R. Ortega, C. Canudas, and S. I. Seleme, Nonlinear control of induction motors: Torque tracking with unknown load disturbance, IEEE Trans. Automat. Contr., vol. 38, pp. 16751680, 1993. G. Espinosa-Perez and R. Ortega, An output feedback globally stable controller for induction motors, IEEE Trans. Automat. Contr., vol. 40, pp. 138143, 1995. G. Espinosa-Perez, R. Ortega, and P. J. Nicklasson, Torque and ux

MARINO et al.: GLOBAL ADAPTIVE OUTPUT FEEDBACK

983

[37] [38]

[39] [40] [41]

tracking of induction motors, in Proc. 3rd European Control Conf., Rome, Italy, 1995, pp. 342347. R. Marino, S. Peresada, and P. Tomei, Output feedback control of current-fed induction motors with unknown rotor resistance, IEEE Trans. Contr. Syst. Technol., vol. 4, pp. 336347, 1996. J. Hu and D. M. Dawson, Adaptive control of induction motor systems despite rotor resistance uncertainty, in Proc. American Control Conf., Seattle, WA, 1995, pp. 13971402; also in Automatica, vol. 32, no. 8, pp. 11271143, 1996. P. C. Krause, Analysis of Electric Machinery. New York: McGrawHill, 1986. J. B. Pomet and L. Praly, Adaptive nonlinear regulation: Estimation from the Lyapunov equation, IEEE Trans. Automat. Contr., vol. 37, pp. 729740, 1992. K. S. Narendra and A. M. Annaswamy, Stable Adaptive Systems. Englewood Cliffs, NJ: Prentice Hall, 1989.

Sergei Peresada was born in Donetsk, Ukraine, on January 14, 1952. He received the Diploma of Electrical Engineer from Donetsk Polytechnical Institute in 1974 and the Candidate of Sciences degree in electrical engineering from the Kiev Polytechnical Institute, Ukraine, 1983. From 1974 to 1977 he was a Research Engineer in the Department of Electrical Engineering, Donetsk Polytechnical Institute. Since 1977 he has been with the Department of Electrical Engineering, Kiev Polytechnical Institute, where he currently is an Associate Professor. From 1985 to 1986 he was a Visiting Professor in the Department of Electrical and Computer Engineering, University of Illinois, Urbana-Champaign. His research interests include applications of modern control theory (nonlinear control, adaptation, VSS control) in electromechanical systems, model development, and control of electrical drives and internal combustion engines.

Riccardo Marino was born in Ferrara, Italy, in 1956. He received the degree in nuclear engineering and the M.S. degree in systems engineering from the University of Rome La Sapienza, in 1979 and in 1981, respectively, and the D.Sc. degree in system science and mathematics from Washington University, St. Louis, MO, in 1982. Since 1984 he has been with the Department of Electronic Engineering at the University of Rome Tor Vergata where he is currently a Professor of systems theory. He is the author, with P. Tomei, of Nonlinear Control Design (Englewood Cliffs, NJ: Prentice-Hall, 1995). His research interests include theory and applications of nonlinear control.

Patrizio Tomei was born in Rome, Italy, on June 21, 1954. He received the dottore degree in electronic engineering in 1980 and the dottore di ricerca degree in 1987, both from the University of Rome La Sapienza. He currently is an Associate Professor of Adaptive Systems at the University of Rome Tor Vergata. He is coauthor of the book Nonlinear Control Design (Englewood Cliffs, NJ: Prentice-Hall, 1995) with R. Marino. His research interests include adaptive control, nonlinear control, robotics, and control of electrical machines.

Anda mungkin juga menyukai