Anda di halaman 1dari 11

Nonlinear Control of Electric Machines: An Overview

David G . Taylor
ue to the increasing availability of improved power electronics and digital processors at reduced costs, there has been a trend to seek higher performance from electric machine systems through the design of more sophisticated control systems software. There exist significant challenges in the search for improved control system designs, however, since (i) the dynamics of most electric machine systems exhibit significant nonlinearities, (ii)not a l l state variables are necessarily measured, and (iii) the parameters of the system can vary significantly from their nominal values. In recent years, a wide range of nonlinear methods for feedback control, state estimation, and parameter identification have emerged, and some of these results are reviewed and summarized in this article. the design of electric machines and for their real-time control, is now contributing more to the level of innovation in electric machine systems than perhaps any other factor. This article attempts to provide an overview of recent developments in control systems design for electric machines operated as motor drives. The article takes a broad perspective in the sense that the wide variety of different machine types is considered, hopefully from a unifying point of view. On the other hand, in order to limit the scope substantially,an effort was made to focus specifically on the more recent nonlinear control methods, as opposed to the classical methods, which have less potential for achieving high dynamic performance. An unavoidable limitation of the presentation is a lack of depth and detail in the descriptions of individual achievements; however, the intention was to highIntroduction Electric machines are electromechanical energy converters, light nonlinear control technology for electric machines to a used for both motor drives and generating systems. Nearly all broad audience, and to guide the interested reader to the approelectric power used throughout the world is generated by so- priate sources for further study. called synchronous machines (operated as generators), and a To put the control design problem into perspective, the article large fraction of all this electric power is consumed by so-called begins with a discussion of electric machine modeling, with induction machines (operated as motors). There are many other types of electric machines, though, distinguished by the materials emphasis on a generic formulation that can be further specialized used, certain key construction features, and the underlying prin- to specific machine types as needed. The most commonly made simplifications available for electric machine models are reciples of operation. The first DC machine was constructed by Faraday around viewed, and the practical issue of model parameterization is also 1820, the first practical version was made by Henry in 1829,and considered. The next topic addressed is (state-feedback) control the first commercially successful version was introduced in 1837. design for electric machines and, for purposes of organization, The three-phase induction machine was invented by Tesla around the recent nonlinear methodologies are divided into three groups: 1887. Although improved materials and manufacturing methods exact linearization design, energy shaping design, and backstepcontinue to refine electric machines, it is safe to say that the ping and manifold design. Most, if not all, of the recently reported fundamental issues relating to electromechanical energy conver- electric machine control contributions can be roughly placed sion have been established for well over a century. within these groups. Finally, the roles of parameter identifiers In such an apparently well-established field, it may come as and state observers for electric machine control are briefly disa surprise that today there is more research and development cussed. activity related to electric machine systems than ever before. The list of references included is by no means exhaustive, and Included in a modern electric machine system is the electric machine itself, power electronic circuits, electrical and/or me- is intended only to represent some key publications for each of chanical sensors, and digital processors equipped with various the topics addressed within the confining structure of this oversoftware algorithms. The recent developments in power semi- view article. Although each of the references cited presumably conductors, digital electronics, and permanent-magnet materials could provide useful background information, two references in have led to enabling technology for todays advanced electric particular deserve special mention. The surveykhallenge paper machine systems. But the increasing use of computers, both for [46], dating back to 1986, not only provides clear descriptions of modeling, estimation, and control issues for electric machines The author is with the Georgia Institute of Technology, School of that are still relevant today, it also contains an extensive bibliogElectrical and Computer Engineering, Atlanta, GA 30332-0250. raphy and would make a helpful supplementto the present article. This work was supported inpart by the National Science Foundation Also, a very recent special issue of Proceedings ofthe IEEE, [lo], under Grant ECS-9158037 and by the Air Force Ofice o f Scientijk contains a collection of informative papers spanning the broad Research under Grant F49620-93-1-0147. areas of power electronics and motion control.
~~ ~ ~

December 1994

0272-1708/94/$04.0001994EEE

41

with a constant current in (l), and that the corresponding component equations in (5) are dropped. General Machine Model Though the model (1)-(5)is not expressed in standard stateAccurate lumped-parameter models for electric machines can space form, certain features worth noting are discernible. There be readily derived from first principles (see [44] formore details). are several nonlinearities to contend with; (1) and (2) enter (5) Although there are many apparently different machine types in and (4) directly, and the load torque 71 could be nonlinear as well. existence, it is worth noting that all electric machines obey the In some cases, several components of the h vector (or the i vector) same fundamental laws of nature. Hence, in order to unify the cannot be measured and, in other cases, it is the mechanical states presentation, this article will not deal with the specialized models (e, o)that cannot be measured. Furthermore, the parameters of specific machine types, but instead will attempt to cover the characterizing the mechanical load (through 71 and J) are almost relevant modeling features of a so-called general machine. The lack of detail is intentional, since this material is covered only to certainly not well known, the resistance R may drift with temthe extent that it will assist the presentation of the main topic, perature changes, and even the parameters defining cp may be subject to uncertainty. The presence of such nonlinearities, tononlinear control design. Consider a general rotating machine with phase windings on gether with the lack of measurements for certain state variables, the stator and rotor (permanent magnets will be considered later, and the probability of parametric uncertainty, all combine to and non-rotating linear motion devices require only minor modi- make the electric machine control system design problem a fications). The angular position of the rotor will be denoted by 8, challenging one. Specific machine types are modeled by assigning the approthe vector of (stator and rotor) phase flux linkages will be denoted priate function cp and, in some cases, assigning zero values to by h, and the vector of (stator and rotor) phase currents will be denoted by i. Application of Amperes law and Gausss law leads certain elements of v (to account for short-circuited rotor windings). As pointed out in [46], there is a potential advantage in to the relationship working with the general machine model if possible, rather than prematurely specializing the model equations to a specific machine type. where cp is a nonlinear function such that, for each 8, i uniquely determines h and vice versa. Moreover, cp is a periodic function of 8 with period 2 d N where N is an integer. The torque produced ,determined by conservation of by the machine, denoted by T ~ is energy to be Model Simplifications Under certain circumstances simplifications can be made to the general machine model, making the control design problem more tractable. One such simplification is valid whenever the excitation to the machine is sufficiently small, such that magnetic saturation in the iron is avoided. Smooth air gap machines are typically designed for operation within the regime of linear magnetics, whereas salient air gap machines almost certainly will encounter magnetic nonlinearities within their intended operating range. For machines that can be considered magnetically linear, the flux linkage relationship (1) may be replaced by the simpler expression

Modeling

Tm

a (a) =-jcp dP
0

Application of Faradays law and Ohms law to the electrical subsystem, and Eulers law to the mechanical subsystem, leads to the dynamic model of the general machine (with constant inertial load and without shaft compliance) summarized by

dh _ -v-Ri dt

where equivalent phases modeling permanent magnets have been included for sake of completeness (im is the equivalent constant current and h m is the equivalent flux linkage). The comparative simplicity of (6) is due to the explicit appearance of a symmetric positive definite inductance matrix relating flux linkage and current. Substituting (6) into (2), the torque produced by the magnetically linear machine may be explicitly written as the sum of three component torques, i.e.

where t denotes time, o is the angular velocity of the rotor, J is the rotor and load inertia, 71 is the load torque (which may depend on time, position, and/or velocity), v is the vector of (stator and rotor) phase voltages, and R is a diagonal positive definite matrix of (stator and rotor) phase resistances. If the electric machine contains permanent magnets, the above formulation is still correct provided that each magnet is modeled as an equivalent phase

The fist torque component (quadratic in i ) is the reluctance torque, the second torque component (linear in i) is the magnetic torque, and the third torque component (independent of i ) is the cogging torque, which has an average value of zero over a period. If the machine does not include magnets, then only the reluctance

42

IEEE Control Systems

torque is present (since im = 0). Needless to say, the dynamic model (3)-(5)presents a more convenient control problem for the special case of magnetically linear machines. A further simplification occurs whenever the inductance matrix of a magnetically linear machine varies sinusoidally with respect to 9. The absence of significant harmonics is characteristic of smooth air gap machines with sinusoidally wound phases, such as the induction machine, the wound-rotor synchronous machine, and the permanent-magnet synchronous machine. On the other hand, for salient air gap machines such as the switched reluctance motor, the stator is constructed with concentrated windings and the variation in inductance (if saturation were neglected) would rarely be sinusoidal. Whenever the harmonics are negligible, the representation of the inductance matrix may be significantly simplified, but this is not really the main benefit. The more significant feature of magnetically linear machines with negligible harmonics is that their models may be simplified through reference frame transformations. Simplifying transformations dating back to the 1920s are applicable to symmetric three-phase machines [59, 601. The purpose of reference frame transformations is to eliminate the explicit dependence of the machine model on the rotor position e. With no dependence on 9, the structure of the model equations is simplified, facilitating both analysis and design. The transformation is itself a &dependent change of coordinates, applied to each of the electrical subsystem variables (v,i, h). The definition of the desired transformation depends on the preferred frame of reference. Common choices are to project all voltages, currents, and flux linkages to a frame of reference that is fixed to the stator, that rotates with the rotor, or that rotates in synchronism with some applied AC excitation. In fact, it is not actually necessary to pre-specify the frame of reference before proceeding with the desired modeling, analysis, or design, due to the notion of the so-called arbitrary reference frame [44]. Efforts have been made to extend the classical reference frame transformations, so as to reach a wider variety of machine types (e.g., see [79,47]). Another approach that can be used to simplify electric machine models, which is valid when the time scales of the electrical and mechanical subsystems are significantly separated, is reduction of the dynamic order of the model. The most common way of dealing with two-time-scale dynamics is through singular perturbation methods [43]. Applications of this approach appear for in [2], for wound-rotor synchronous generators, and in [a], permanent-magnet synchronous motors (from a sampled-data perspective). An altemative approach, selective modal analysis, has been used in [73] to study the mid-frequency resonance phenomena associated with open-loop constant-frequency excitation of hybrid permanent-magnet stepper motors. Whenever order reduction can be achieved without introducing excessive approximation error, both analysis and design are simplified by virtue of improved numerical conditioning, lower dimensionality, and lower bandwidth. Model Parameterization Along with the opportunities provided by computer-aided modeling and microprocessor-based control comes the potential to select accurate yet efficient model parameterizations. The electric machine model must be accurate enough to capture all the essential characteristics relevant to control design, but with-

out being so detailed that the complexity of control implementation is prohibitive. Another desired feature of the model parameterization would be that it be compatible with a reliable method for parameter identification. This issue is best illustrated by considering the primary nonlinearities which may appear in (1). The nonlinearity of h with respect to 8 is due to the spatial periodicity of any rotating machine. Consequently, it would seem natural to capture this nonlinearity with a truncated Fourier series. The nonlinearity of h with respect to i is due to magnetic saturation effects. Hence, exponential or hyperbolic tangent functions would seem the logical choice for approximating this nonlinearity. On the other hand, even if such physically motivated parameterizations prove to be accurate, the real-time computational requirements for any corresponding model based controller could prove troublesome. A reasonable altemative would be to represent the nonlinearities in piecewise fashion, using a unique multivariate polynomial within each individual domain partitioning the operating space of interest. For example, consider a switched reluctance motor with uncoupled phases. In this case, the flux linkage h of a given phase depends only on the current i of that same phase and the rotor position 8. In each domain of the (i, 0) plane, the flux linkage relationship (1) for a given phase may be represented by

where (9)

and where C is a domain-dependent coefficient matrix. At domain boundaries, where distinct coefficient matrices serve to describe the flux linkage, continuity and a degree of smoothness can be imposed. The choice of cubic polynomials in this example is arbitrary. An excellent development along these lines for switched reluctance motors appears in [51] and, to a lesser extent, a piecewise approach for modeling permanent-magnet motors exhibiting saturation and reluctance variation is presented in [32]. In [a], a piecewise-linear approach to modeling the nonlinearity of a permanent-magnet synchronous motor was developed, and the resulting computationally efficient implementations of both self-tuning and sensorless control were presented. The benefits of piecewise-polynomial modeling of nonlinearities include: suitability to least squares fitting techniques for evaluation of polynomial coefficients from measured data; ease of computation, e.g. torque coefficients are easily computed from flux linkage coefficients, and control computations often require simply finding the roots of a polynomial; economy of storage, in the sense that many fewer polynomial coefficients need to be stored compared to the size of an entire data set characterizing the nonlinearity values at grid points.

December 1994

43

I - ! -

I I

Prior to the development of a formal theory for exact linearization design, closely related nonlinear feedback control schemes had already been developed for the induction motor. Exact Linearization Design Some of the most common control designs for electric ma- The classical field oriented control, introduced in [5] over chines today, for applications requiring high dynamic perform- twenty years ago, involves the transformation of electrical variance, are based on forms of exact linearization [36]. The design ables into a frame of reference which rotates with the rotor flux concept is reflected by the two-loop structure of the controller: vector (the dq frame). This reference frame transformation, toin the first design step, nonlinear compensation is sought which gether with a nonlinear feedback, serves to reduce the complexity explicitly cancels the nonlinearities present in the motor (without of the dynamic equations, provided that the rotor flux is not regard to any specific control objective), and this nonlinear identically zero. Under this one restriction, the rotor flux amplicompensation is implemented as an inner feedback loop; in the tude dynamics are made linear and decoupled and, moreover, if second design step, linear compensation is derived on the basis the rotor flux amplitude is regulated to a constant value, the speed of the resulting linear dynamics of the pre-compensated motor dynamics will also become linear and decoupled. Provided that to achieve some particular control objective, and this linear the rotor flux amplitude may be kept constant, the field oriented compensation is implemented as an outer feedback loop. The control thus achieves an asymptotic linearization and decouadvantage of linear closed-loop dynamics is clearly that selection pling, where the d-axis voltage controls rotor flux amplitude and of controller parameters is simplified, and the achievable tran- the q-axis voltage controls speed. Although the field oriented approach to induction motor sient responses are very predictable. control is widely used today and has achieved considerable Not all nonlinear systems can be controlled in this fashion; the applicability of exact linearization is determined by the type success, the formal use of exact linearization design can provide and location of the model nonlinearities. Furthermore, exact altemative nonlinear control systems of comparable complexity, linearization is not really a single methodology, but instead but achieving true (as opposed to asymptotic) linearization and represents two distinct notions of linearizability, though in both decoupling of flux and torque or speed. For example, using a cases the implementation requires full state feedback. In the first reduced-order electrical model (under the assumption that the rotor speed is constant), an input-state exact linearization static case, exact linearization of the input-output dynamics of a system state-feedback design is reported in [23] that achieves complete is desired, with the output taken to be the controlled variables. decoupling of the rotor flux amplitude and torque responses. The This case, referred to as input-output linearization, is the more full-order electromechanical model of an induction motor tums intuitive form of exact linearization, but can be applied in a out not to be input-state exactly linearizable [52]. However, as straightforward way only to so-called minimum phase systems shown in [45] (see also [52]), input-output exact linearization (those systems with stable zero dynamics). A system can be methods do apply to the full-order electromechanical model. input-output linearized if it has a well-defined relative degree The input-output exact linearization control for induction (see [36] for clarification). In the second case, exact linearization motors in [45, 521 begins with an appropriate description of the of the entire state-space dynamics of a system is desired, and no model. Using standard notation from the theory of exact linearioutput needs to be declared. This case, referred to as input-state zation, the induction motor model can be put into the form linearization, has the advantage of eliminating any potential difficulties with intemal dynamics but is less intuitively appealing and can be more difficult to apply in practice. Input-state linearization applies only to systems that are characterized by integrable, or nonsingular and involutive, distributions (see [36] for clarification). Standard models of most electric machines are exactly linearizable, in the sense(s) described above. Prior literature has disclosed many examples of exact linearization applied to electric machines, including experimental implementation, for various machine types and various types of models. For any machine where x is the state vector, ul and u2 are the scalar control inputs, type, the most significant distinction in the application of exact and yi and y2 are the scalar outputs to be controlled. The compolinearization relates to the order of the model used. When full- nents of the state vector (in a fixed stator frame) are the rotor order models are used, the stator voltages are considered to be speed xi, the q-axis rotor flux x2, the d-axis rotor flux x3, the the control inputs. When a reduced-order model is used, the q-axis stator currentxq, and the d-axis stator currentxj. The inputs assignment of the control inputs depends on how the order are the q-axis and d-axis stator voltages, ul and u2, and the outputs reduction has been performed: if the winding inductance is are the rotor speed yl and the square of the rotor flux amplitude neglected, then voltage will still be the input but the mechanical y2. With these assignments, the state equation (1 1) is completely specified by subsystem model will be altered; if a high-gain current loop is employed, then current will be the input in an unaltered mechanical subsystem model. In either case, exact linearization provides a systematic method for designing the electronic commutation A4 = functions within the inner nonlinear loop, so that the outer linear loop is concemed only with the motion control part of the total system.

Control Design Methods

44

IEEE Control Systems

reduced from nominal as the speed reference is increased above rated speed, in order to keep the required feed voltages within the inverter limits. Operation in this flux-weakening regime will excite the coupling between flux and speed in the classical field g2(x) = [ 0 0 0 0 d IT (16) oriented control, causing undesired speed fluctuations (and perhaps instability). where constants mi, m2, ai,a2, bi, b2, c, and dare related to the There are other motivations for considering time-varying flux parameters of the motor and load (see [451 or [52] for modeling references as well. For instance, in [40,41] the flux is adjusted details). The output functions in (12) and (1 3) are given by as a function of speed in order to maximize power efficiency (i.e. only the minimum stator input power needed to operate at the desired speed is actually sourced). In [40], the flux reference is computed on the basis of predetermined relationships derived off-line and the control is implemented in a reference frame h2(x) = xE + xs (18) rotating with the stator excitation. In [41], the flux reference is computed on-line using a minimum power search method and With rotor flux amplitude and speed chosen as outputs to be the control is implemented in a fixed stator frame of reference. controlled, simple calculations show that the system has well-de- The earlier work of [42] illustrates an effort to optimize efficiency fined relative degree (2, 2), provided that the rotor flux is not using the classical field-orientation concepts. Yet another possiidentically zero. In other words, the system of equations bility, presented in [7,8], would be to vary the flux reference as a function of speed in order to achieve optimum torque (maximum for acceleration and minimum for deceleration) given limits on allowable voltage and current. For comparison, see the earlier work [77J which considers the same torque optimization objeccan be solved for ui and u2 (where vi and v2 are outer-loop tive using field-orientation. In each of these references, [40,41, controls), by virtue of the invertibility of the square matrix 7, 81, high-gain current loops are used so that the exact lineariappearing on the left-hand side of (19), provided that the rotor zation is performed with respect to current inputs rather than flux is not identically zero. The nonlinear functions appearing in voltage inputs. Clearly, exact linearization permits optimization the inner-loop nonlinear compensation(19), which are simple to goals (which require variable flux references) and high dynamic calculate given (14)-(18), are so-called Lie derivatives of the performance in position, speed, or torque control, to be achieved output functions (see [36] for clarification). If the nonlinear simultaneously. Exact linearization has been suggested for the control of many feedback (19) is applied to system (1 l), then the two outputs yi and y2 will have linear and decoupled input-output behavior other types of electric machines as well. Both input-state exact linearization and input-output exact linearization are used to modeled by design controllers for wound-field bmsh-commutated DC motors, in [55, 18, 191. For various permanent-magnet machines, there are many references illustrating the use of exact linearization. For instance, in [31], a three-phase wye-connected permanent-magnet synchronous motor with sinusoidally distributed windings is modeled with piecewise-constantparameters (which depend on current to account for reluctance variations and magnetic saturation), and an input-state exact linearization controller inputwhere s is the Laplace variable. Linear outer-loop compensation is derived from the rotor reference frame model. In [80], state exact linearization is applied to the hybrid permanent-magcan easily be. designed for yi andy2 on the basis of (20) and (21). Hence, under the constraint of nonzero rotor flux, it is possible net stepper motor with cogging torque accounted for, and it is to derive a nonlinear static state-feedback that controls rotor flux further shown how constant load torques may be rejected using amplitude and speed in a noninteracting fashion, with linear a nonlinear observer. This work was continued in (91, where the second-order transients for each controlled output (and with experimental implementation is described, including treatment bounded first-order internal dynamics [52]). Although the full- of practical issues such as speed estimation from position sensors order induction motor model is not input-state exactly and operation at high speeds despite voltage limitations.Whereas linearizable, the augmented system obtained by adding an inte- each of the above references considers the full-order electromeinvolves exact linearization grator to one of the inputs does satisfy this property locally (but chanical model, the approach of [64] only for speed control rather than position control, and only by of permanent-magnetsynchronous motors using a reduced-order discrete-time model suitable for direct digital control, accounting using a complicated nonlinear controller) [171. Although the difference between the classical control of [5] fully for harmonic effects. Finally, considering switched relucand the exact linearization control of [45,52] may appear to be tance motors, exact linearization has been applied to the full-orand the reduced-order model (assuming current a minor one, the complete decoupling of speedand flux dynamics der model in [35] in the closed-loop system (during transients as well as in steady tracking) in 1751, in both cases accounting for harmonic effects state) provides the opportunity to optimize performance. For and magnetic saturation. It is worth pointing out that when the example, as mentioned in [521, the flux reference will need to be models rotor position dependence is not purely sinusoidal, refg1(x)=[ 0 0 0 d 0

IT

December 1994

45

erence frame transformations have not typically been used for control design. Optimization objectives can be considered within exact linearization designs for these other types of machines, too. For example, [76] attempts to reduce the peak values and the ratesof-change of reference currents in reduced-order switched reluctance motor controllers based on exact linearization. In [28], a formal numerical optimization approach to the same problem is taken, but with goals of minimizing copper losses or feed voltages. The exact linearization control for permanent-magnet synchronous motors in [64], discussed above, is selected to minimize copper losses. One of the most interesting formulations of the optimization problem is reported in [29], in the context of brushless permanent-magnet motors without reluctance variation or magnetic saturation effects. Under the assumption of high-gain current tracking, the reduced-order exact linearization problem is parameterized such that all possible instantaneous torque controllers are characterized by the free choice of a single periodic function. The problem of optimizing some desired feature (such as efficiency or feed voltage) while assuring instantaneous torque control then reduces to the optimal selection of this free function. To illustrate typical performance characteristics of a motion control system designed using exact linearization, an experimental testbed developed at Georgia Tech was configured with a switched reluctance motor. a frictional load, and a two-quadrant PWM current-tracking amplifier. The motor has 6 stator poles and 8 rotor poles, a phase resistance of 5 R, and a phase inductance that varies between I O mH and 50 mH depending on the rotor position. Rotor position was measured using a 768 line optical encoder, and rotor velocity was obtained as the slope of a straight line least-squares fit through four position measurements. All control software was executed on a Texas Instruments TMS320C30 digital signal processor, with a sampling interval of 1 msec. The motor, load, and encoder are shown in Fig. 1. Current-tracking exact linearization control requires knowledge of the motor's phase torque-angle characteristic over a range of phase currents. Given some desired torque, command currents for each phase are selected such that the total torque produced is independent of the rotor position. For a three-phase motor, these command currents are uniquely determined by the desired torque at rotor positions for which only one phase can produce torque of the desired polarity. At other rotor positions for which two phases may contribute, the command currents may be selected to achieve optimization objectives while still guaranteeing production of the desired torque. For the present experiment, the command currents were computed according to the method discussed in [76]. The motor was commanded to smoothly transition between rotor positions two revolutions apart, with transition times of 0.25 sec. Shown in Fig. 2 is the response of the closed-loop system, as well as the command currents during a short time interval. Even though this system has not been carefully tuned, the peak position errors during transitions are less than the step angle of 15". The steady-state positioning accuracy of the system is better, with steady-state offsets of only about 1 encoder count or 0.1'. The command currents reveal the sequential nature of the phase excitation, with the shape of a given excitation pulse determined by the nonlinear inner-loop compensation which receives the

___ Fig. I . The experimental mitched reluctance motor with load and encoder
~

torque command from a linear outer-loop compensation based on motion tracking errors. Energy Shaping Design A new approach to control design for electric machines, referred to as energy shaping, is based on the Euler-Lagrange form of the dynamic equations (as opposed to the state-space form). Although the approach applies in principle to all electric machines, it has first been applied to the problem of torque regulation in induction motors [56, 571, using the full-order electromechanical model. A motivation for this approach comes from the literature on control of mechanical systems (e.g., robot manipulators, high-performance aircraft, underwater and space vehicles), as discussed in [67]. In contrast to the exact linearization design method, which could arguably be characterized as a mathematically motivated approach, the energy shaping design approach has evolved from consideration of the physical properties of the system, such as energy conservation and passivity. When applied to the mechani-

46

IEEE Control Systems

Time (sec)

Time (sec)

Fig. 2. Position control response of switched reluctance motor using current-trackingexact linearization design,

cal robot manipulator, energy shaping design has led to numerous advantages over exact linearization design (see [67] for details). For instance, the needs for on-line inversion of an inertia matrix, for measurement of joint accelerations, and for use of projection to restrict parameter updates in adaptive implementations, are all eliminated. Another distinguishing feature is that, when energy shaping design is applied to a nonlinear system, the closed-loop dynamics remain nonlinear. A key step for energy shaping design is to express a certain matrix, associated with the Euler-Lagrange equations, in a special skew-symmetricform. This representation is not mathematically motivated, but instead is simply a way of exhibiting the systems workless forces, i.e. the terms that do not influence the energy balance. Due to the fact that these particular terms do not affect the stability properties of the system, there is no need to cancel or offset them with feedback control, thus simplifying somewhat the control design and enhancing robustness. Unlike the models of many mechanical systems, the EulerLagrange equations for electric machines have fewer control inputs than degrees of freedom. This fundamental difference, which complicates the design problem considerably, is accounted for in the extension of [67] found in [56]. In Euler-Lagrange form, the so-called generalized coordinates q of the electric machine are the electric charges (rather than currents or flux linkages) and the rotor position, but (at least for the induction motor) 4 does not explicitly appear in the model. Other differ-

ences between the electric machine and mechanical system versions of energy shaping design are that, for electric machines, it is 4 that is to be controlled (rather than q) and, more significantly, dynamics must be introduced into the state-feedback control algorithm (Le. static state-feedback is not sufficient). There are some significant advantages of energy shaping design over exact linearization. As a consequence of being based on physical energy properties of the system, the energy shaping design tends to be robust with respect to modeling errors, provided that the perturbations do not violate the skew-symmetry assumption of the model. The use of the Euler-Lagrange equations also facilitates adaptive design, since the key features of linear and minimal parameterization are sure to be present, whereas these features are often lost using state-space formulations. For electromechanical systems which can be viewed as feedback interconnections of passive electrical and mechanical subsystems, energy shaping doesnt have to be applied to the overall system. Instead, the option exists to apply energy shaping only to the electrical subsystem and then treat the mechanical subsystem as a passive perturbation. As shown in [26], this subsystem partitioned approach can be applied to induction motors without requiring a rotor flux observer (or rotor flux measurement). In addition to these generic benefits, the energy shaping design for torque regulation in induction motors [57] results in a globally defied control law which is free of singularities, permitting operation from start-up without the need to pre-magnetize the rotor, and this control law achieves the torque regulation objective globally, so that initial conditions are not restricted. In [57], it is also shown how to use energy shaping ideas to design a globally convergent observer for the (possibly unmeasured) rotor electrical variables, as well as a globally convergent adaptive version of the control to compensate for unknown constant values of rotor resistance and/or load torque (but under the assumption of full state measurement). The formal proof of the globally stable (and globally defined) observer-based control of [57] deserves particular emphasis, as this result appears to be the first of its kind. Some restrictions in the results of [57] have been removed in [58], so that observer-based torque tracking with unknown but linearly parameterizedload torque is made possible. It is also worth mentioning that the energy dissipation properties of the system are not affected by nonsingular state coordinate changes, so that the energy shaping controller could be re-designed using a different frame of reference than in [57,58], if desired. For instance, in [25, 271 the energy shaping idea is applied directly in a fixed stator frame (instead of using the reference frame rotating synchronously with the applied sinusoidal voltage, as in [57, 581) to design observer-based control, thereby eliminating the need for on-line implementation of any reference frame transformations to implement the control. On the other hand it appears, at least at this early stage of development, that the energy shaping approach applied to electric machines has several disadvantages as well. First and foremost, in its present form, the torque regulatiodtracking energy shaping designs discussed above would require measurement of rotor acceleration if position or velocity control loops were used to feed the torque control loop. More recent work [26] appears to be addressing this drawback, though, by effectively using an open-loop assignment of the desired torque command signal in terms of the desired velocity trajectory. A more minor limitation,

December 1994

47

due to the fact that the Euler-Lagrange equations of electric machines have fewer control inputs than degrees of freedom, is that dynamics must necessarily be included in the controller even if neither state observation nor parameter identification is needed. Another concem (at the present time) is simply the lack of documented applications of energy shaping results for electric machines other than the induction motor, together with comparisons to existing control approaches for these other electric machines. Finally, no experimental work on energy shaping control has yet been reported. It would seem premature to conclude at this point that energy shaping offers significant advantages to all machines in all applications; however, it should be clear that research along these lines has potential and, hopefully, further results will be forthcoming. Backstepping and Manifold Designs Yet another alternative to exact linearization design for electric machines has been recently developed, based on integrator backstepping and nonlinear damping. These are among several nonlinear design tools which make up the design toolkit of [38]. Such tools can be used in a systematic fashion, as shown in [38], and can be applied to systems that are not exactly linearizable as well as to systems possessing disturbances. The justification for these design tools is established through Lyapunov analysis. The basic idea of backstepping design is to select an appropriate function of state variables to be a pseudo-control input for a lower dimensionsubsystemof the overall system. The selection is made so as to simplify the first stage of design, in the sense that it should be more or less obvious how to design the pseudocontrol feedback if indeed there were no other dynamicspresent. The pseudo-control is of course not a true control input so, to account for this fact, an error variable representing the deviation of the pseudo-control from its desired value must be introduced. When the dynamics of this pseudo-control error variable are written, the true control input may still not appear, in which case the whole procedure is repeated. Each backstepping stage results in a new pseudo-control design, expressed in terms of the pseudocontrol designs from preceding design stages. Finally the procedure terminates, resulting in a feedback design for the true control input which achieves the original design objective by virtue of a final Lyapunov function, formed by summing up the Lyapunov functions associated with each individual design stage. This backstepping design formulation was first applied to electric machines in [39], and to electrically-driven mechanical manipulators in [21]. In [39], a backstepping approach is used to design a nonlinear feedback control for induction motors, without requiring flux measurements, but under the assumption that all the parameters of the system are known. A pseudo-control proportional to the q-axis stator current is selected for the torque control objective, and a pseudo-control proportional to the d-axis stator current is selected for the flux control objective. To terminate with a design for the stator voltages, one stage of backstepping is needed for both inputs, and decoupling is achieved within the region that a certain decoupling matrix is nonsingular. Though the resulting controller is not globally defined and does not provide speed and flux regulation from all initial conditions, it has the advantages of guaranteed local stability (using a flux observer) and reduced sensitivity to flux estimation errors (in comparison with the exact linearization approach).

In [21], the emphasis is on a multiple degree of freedom mechanical subsystem, with the electrical subsystem dynamics chosen to model permanent-magnet brush-commutated DC motors. The backstepping approach is used to design state feedback nonlinear control, first under an assumption of known parameters, and then with the possibility of parametric uncertainty and bounded disturbances included. The pseudo-control selected for position trajectory control is the armature current, and one stage of backstepping is needed to reach the final armature voltage control. Despite the simplicity of the electrical subsystem in [21], it is not difficult to extend the design for other classes of electric machines. In fact, state-feedback robust and adaptive backstepping designs for induction motors [22, 341, for permanent-magnet synchronous motors [13,33], for hybrid permanent-magnet stepper motors [68,69], and for switched reluctance motors [12, 141 are all available. Note that in all the cited applications to electric machines, the backstepping procedure boils down to the selection of current as a pseudo-control, with voltage being considered the true control input. The motivation for temporarily considering state-dependent functions to be pseudo-controls, at least from the point of view of integrator backstepping methodology, is primarily mathematical rather than physical in nature. However, there is an alternative point of view motivated by time-scale considerations which can lead to similar design concepts. For instance, if the voltage-to-current transients are quite fast compared to the mechanical transients of greatest interest, then it makes good engineering sense to consider neglecting such transients altogether (provided that they are stable). This would intuitively lead to reduced-order control designs, with either voltages or currents as inputs. The design concept suggested above can be referred to as a slow manifold design (sometimes called composite control), and is formally covered in the singular perturbation literature (see [43] for an extensive development). A closely related design concept is sliding-mode control. As applied to electric machines, the two-time-scale approach to control design essentially arises in either of two ways. In the first case, the time-scale separation is due to the presence of a negligible stator inductance whereas, in the second case, it is due instead to the use of a high-gain current loop. Arefinement of standard composite control makes use of correction terms, based on integral manifold conditions, in order to guarantee that the design objective will be asymptotically met despite the possibly non-negligible value of the singular perturbation parameter. For general systems, this correction can be done only approximately (but to any desired order in the singular perturbation parameter) but, for electric machine systems, a single correction term that can be computed in closedform suffices to provide exact asymptotic performance. This point is made in [63, 721 and, although the objective in both references relates to exact linearization of the mechanical subsystem, in fact the correction term is generic and can be used to achieve exact asymptotic performance from any mechanical subsystemdesign. It is interesting to note that the two-time-scale reduced-order corrective design of [63,72] is very closely related to both the energy shaping designs and the backstepping designs, when applied to electric machines, though this point has not yet been made clear in the literature. Both backstepping design and manifold design can be viewed as extensions to simple reduced-order exact linearization meth-

48

IEEE Control Systems

r-

T1

ods. However, while the latter has been thoroughly studied on the experimental level, the backstepping and manifold designs have yet to be closely examined in the experimental sense.

proach already mentioned is based on energy shaping. A thorough discussion of this estimation problem, with a complete historical perspective, is given in [74].

Concluding Remarks

References
[l] P.P. Acamley, R.J. Hill, and C.W. Hooper, Detection of rotor position in stepping and switched motors by monitoring of current waveforms, IEEE Transactions on Industrial Electronics, vol. 32, no. 3, pp. 215-222, 1985.

Although the three types of nonlinear control design summarized above reasonably cover the recent research efforts, there are still many supplementary issues that are important to mention. It is often necessary to augment the basic controllers with algorithms capable of providing state estimates or of identifying parameters. Since model based nonlinear controllers depend on parameters which may be unknown or slowly varying, an on-line parameter identification scheme can be included to achieve indirect adaptive control. There are several motives for agreeing to this additional complexity, including the potential for improved reliability due to the use of diagnostic parameter checks. Examples of parameter identification schemes and adaptive control (not already mentioned) may be found in [52,70,53,4] for induction motors, in [6, 61, 62, 15, 651 for various permanent-magnet brushless motors, and in [3] for switched reluctance motors. One of the challenges in practical nonlinear control design for electric machines is to overcome the need for full state measurement. The simplest example of this is commutation sensor elimination (e.g. elimination of Hall-effect devices from inside the motor frame). For electronically commutated motors, such as stepper motors and brushless DC motors, some applications do not require control of instantaneous torque, but can get by with simple commutation excitation at a variable firing angle. Even at this level of control, the need for a commutation sensor is crucial in order to maintain synchronism. The works of [ 11 for permanent-magnet motors and [50] for switched reluctance motors are representative of schemes used to drive commutation controllers. A more difficult problem in the same direction is high accuracy position estimation suitable for eliminating the high resolution position sensor required by most of the nonlinear controls discussed earlier. In some applications, despite the need for high dynamic performance, the use of a traditional position sensor is considered undesirable due to cost, the volume andor weight of the sensor, or the potential unreliability of the sensor in harsh environments. In this situation, the only alternative is to extract position (andor speed) information from the available electrical terminal measurements. Of course, this is not an easy-thingto do, precisely because of the nonlinearities involved. Work done in this area appears in [37,24,66] for permanent-magnet synchronous motors, in [49,30] for switched reluctance motors, and in [71,78,54] for induction motors. A more pervasive state estimation problem concerns the estimation of velocity given a position measurement from a traditional position sensor. One set of approaches, not model based, is reported in [111. The methods considered include linesper-period, reciprocal-time, Taylor series expansions, backward difference expansions, and least-squares curve fits. Yet another approach based on phase-locked loops is presented in [20]. Linear and nonlinear observer-based schemes for velocity estimation from position measurements appear in [48] and [161, respectively. For the induction motor specifically,there is aneed to estimate either the rotor fluxes or rotor currents, in order to implenient many of the nonlinear controllers discussed earlier. One ap-

[2] S. Ahmed-Zaid, P.W. Sauer, M.A. Pai, and M.K. Sarioglu,Reducedorder modeling of synchronous machines using singular perturbation, IEEE l?ansacriuns on Circuits and Systems, vol. 29, no. 11, pp. 782-786, 1982. [3] L.B. Amor, L.A. Dessaint, 0.Akhrif, andG. Olivier, Adaptive feedback linearization for position control of a switched reluctance motor: Analysis and simulation, International Journal o f Adaptive Control and Signal Processing, vol. 7, pp. 117-136, 1993. [4] D.J. Atkinson, P.P. Acamley, and J.W. Finch, Observers for induction motor state and parameter estimation, IEEE Transactions on Industry Applications. vol. 27, no. 6, pp. 1119-1127, 1991.

[5] E Blaschke, The principle of field orientation applied to the new transvector closed-loop control system for rotating field machines, Siemens Review, vol. 39, pp. 217-220, 1972.
[6] A.J. Blauch, M. Bodson, and J. Chiasson, High-speed parameter estimation of stepper motors, IEEE Transactions on Control Systems Technology, vol. 1, no. 4, pp. 270-279, 1993. [7] M. Bodsou and J.N. Chiasson, A systematic approach to selecting optimal flux references in induction motors, Record o f the 1992 IEEE Industry Applications Society Annual Meeting, Houston, TX, pp. 531-537, 1992. [8] M. Bodson, J. Chiasson, and R. Novotnak, High-performance induction motor control via input-output linearization,IEEE Control Systems Ma@zine, vol. 14, no. 4, pp. 25-33, 1994. [9] M. Bodson, J.N. Chiasson, R.T. Novomak, and R.B. Rekowski, Highperformance nonlinear feedback control of a permanent magnet stepper motor, IEEE Transactions on Control Systems Technology, vol. 1, no. 1, pp. 5-14, 1993. [lo] B.K. Bose (editor), Special Issue on Power Electronics and Motion Control, Proceedings o f the IEEE, vol. 82, no. 8, 1994.
[ 111 R.H. Brown, S.C. Schneider, and M.G. Mulligan, Analysis of algorithms for velocity estimation from discreteposition versus time data,IEEE Transactions on Industrial Electronics, vol. 39, no. 1, pp. 11-19, 1992.

[I21 J.J. Carroll, D.M. Dawson, and 2. Qu, Robust tracking control of a switched reluctancemotor tuming an inertial load, Proceedings of the 1992 American Control Conference, Chicago, IL, pp. 1520-1522, 1992.
[13] J.J. Carroll and D.M. Dawson, Robust tracking control of a brushless dc motor with application to direct-drive robotics, Proceedings of the IEEE International Conference on Robotics and Automation, Atlanta, GA, pp. 94-99,1993. [14] J.J. Carroll and D.M. Dawson, Adaptive tracking control of a switched reluctancemotor turning an inertial load, Proceedings o f the 1993American Control Conference, San Francisco, CA, pp. 670-674, 1993.
[15] D. Chen and B. Paden, Adaptive linearization of hybrid step motors: Stability analysis, IEEE Transactions on Automatic Control, vol. 38, no. 6, pp. 874-887, 1993.
[ 161 J.N. Chiasson and R.T. Novotnak, Nonlinear speed observer for the PM stepper motor, IEEE Transactions on Automatic Control, vol. 38, no. 10, pp. 1584-1588,1993.

December 1994

49

lr-

] I -

[17] J. Chiasson, Dynamic feedback linearization of the induction motor, IEEE Transactions on Automatic Control, vol. 38, no. 10, pp. 1588-1594, 1993.
[181J. Chiasson and M. Bodson, Nonlinear control of a shunt DC motor,

[36] A. Isidori, Nonlinear Control Systems, 2nd Edition. New York, NY Springer-Verlag, 1989. [37] L.A. Jones and J.H. h g , A state observer for the permanent-magnet synchronous motor, IEEE Transactions on Industrial Electronics, vol. 36, no. 3, pp. 374-382, 1989. [38] I. Kanellakopoulos, P.V. Kokotovic, and A.S. Morse, A toolkit for nonlinear feedback design, Systems and Control Letters, vol. 18, pp. 83-92, 1992. [39] I. Kanellakopoulos, P.T. Krein, and E Disilvestro, Nonlinear flux-observer-based controlof induction motors, Proceedingsof the I992 American Control Conference, Chicago, IL, pp. 1700-1704,1992. [ a ] D.I. Kim, I.J. Ha, and M.S. KO, Control of induction motors via feedback linearization with input-output decoupling, International Journal of Control, vol. 51, no. 4, pp. 863-883, 1990. [41] G.S. Kim, I.J. Ha, and M.S. KO,Control of induction motors for both high dynamic performance and high power efficiency, IEEE Transactions on Industrial Electronics,vol. 39, no. 4, pp. 323-333, 1992. [42] D.S. Kirschen, D.W. Novotny, and T.A. Lipo, Optimal efficiency control of an induction motor drive, ZEEE Transactions on Energy Conversion, vol. 2, no. 1, pp. 70-75, 1987. [43] P.V. Kokotovic, H.K. Khalil and J. OReilly, Singular Perturbation Methods in Control: Analysis and Design. New York, Ny: Academic Press, 1986. [44] P.C. Krause, Analysis of Electric Machinery. New York, NY McGrawHill, 1986. [45] Z. Krzeminski, Nonlinear control of induction motor, Proceedings of the loth IFAC World Congress, Munich, Germany, pp. 349-354, 1987. [46] J.H. Lang, G.C. Verghese, and M. Ilic-Spong, Opportunities in estima-

IEEE Transactions on Automatic Control, vol. 38, no. 11, pp. 1662-1666, 1993.
[ 191J. Chiasson, Nonlinear differential-geometric techniques for control of

a series DC motor, IEEE Transactions on Control Systems Technology, vol. 2, no. 1, pp. 35-42, 1994. [20] C.F. Christiansen, R. Battaiotto, D. Femandez, and E. Tacconi, Digital measurement of angular velocity for speed control, ZEEE Transactions on Industrial Electronics, vol. 36, no. 1, pp. 79-83, 1989. [21] D.M. Dawson, 2. Qu, and J.J. Carroll, Tracking control of rigid-link electrically-driven robot manipulators, International Journal of Conml, vol. 56, no. 5, pp. 991-1006, 1992. [22] D. Dawson, Z. Qu. and J. Hu, Robust tracking control of an induction motor, Proceedings of the 1993 American Control Conference, San Francisco, CA, pp. 648-652, 1993. [23] A. De Luca and G.Ulivi, Design of an exact nonlinear controller for induction motors, IEEE Transactions on Automatic Control,vol. 34, no. 12, pp. 1304-1307,1989. [24] N. Ertugrul and P. Acarnley, A new algorithm for sensorless operation of permanent magnet motors, IEEE Transactions on Industry Applications, vol. 30, no. 1, pp. 126-133, 1994. [25] G. Espinosa and R. Ortega, Control of induction motor models in a fixed reference frame, presented at the 32nd IEEE Conference on Decision and Control, San Antonio, TX, 1993 (not in proceedings). [26] G. Espinosa and R. Ortega, State observers are unnecessary for induction motor control, to appear in Systems and Control Letters.

[47] X.Z. Liu, G.C. Verghese, J.H. Lang, and M.K. Onder, Generalizing the Blondel-Park transformation of electrical machines: Necessary and sufficient and control [28] Filicori, C.G. Lo Bianco, and A. Tonielli, conditions, IEEE Transactions on Circuits and Systems, vol. 36, no. 8, pp. strategies for a variable reluctance direct-drive motor, IEEE Transactions 1058-1067, 1989. on Industrial Electronics,vol. 40, no. 1, pp. 105-115, 1993. [29] I.J. Ha and C.I. Kang, Explicit characterization of all feedback linearizing controllers for a general t y p e brushless dc motor, IEEE Transactions on Automatic Control, vol. 39, no. 3, pp. 673-677, 1994. [30] W.D. Harris and J.H. Lang, A simple motion estimator for variable-reluctance motors, IEEE Transactions on Industry Applications, vo1. 26, no. 2, pp. 237-243,1990. [31] N. Hemati, J.S. Thorp and M.C. Leu, Robust nonlinear control of brushless dc motors for direct-drive robotic applications, IEEE Transactions on Industrial Electronics,vol. 37, no. 6, pp. 460-468, 1990. [32] N. Hemati and M.C. Leu, A complete model characterization of brushless dc motors, IEEE Transactions on Industry Applications, vol. 28, no. 1, pp. 172-180, 1992. [33]J. Hu, D.M. Dawson, and J.J. Carroll, Anadaptiveintegrator backstepping tracking controller for brushless dc motor robotic load, Proceedingsof the I994 American Control Conference, Baltimore, MD, pp. 1401-1405, 1994. [34] J. Hu, D.M. Dawson, and 2. Qu, Adaptive tracking control of an induction motor with robustness to parametric uncertainty, IEE Proceedings: Electric Power Applications, vol. 141, no. 2, pp. 85-94, 1994. [35] M. Ilic-Spong, R. Marino, S.M. Peresada, and D.G. Taylor, Feedback linearizing control of switched reluctance motors, IEEE Transactions on Automatic Control, vol. 32, no. 5, pp. 371-379, 1987. [48] R.D. Lorenz and K.W. Van Patten, High-resolution velocity estimation for all-digital ac servo drives, IEEE Transactions on Industry Applications, vol. 27, no. 4, pp. 701-705, 1991. [49] A. Lumsdaine and J.H. Lang, State observers for variable-reluctance motors, IEEE Transactions on Industrial Electronics, vol. 37, no. 2, pp. 133-142,1990. [50] S.R.MacMinn, W.J. Rzesos, P.M. Szczesny, and T.M. Jahns, Application of sensor integration techniques to switched reluctance motor drives, IEEE Transactions on Industry Applications, vol. 28, no. 6, pp. 1339-1344, 1992. [51] D.G. Manzer, M. Varghese, and J.S. Thorp, Variable reluctance motor characterization, IEEE Transactions on Industrial Electronics, vol. 36, no. 1, pp. 56-63, 1989. [52] R. Marino, S. Peresada, and P. Valigi, Adaptive input-output linearizing control of induction motors, IEEE Transactions on Automatic Control, vol. 38, no. 2, pp. 208-221, 1993. [53] R. Nilsen and M.P. Kazmierkowski, Reduced-order observer with parameter adaptation for fast rotor flux estimation in induction machines, Proceedings of the IEE, pt. D, vol. 136, no. 1, pp. 35-43, 1989. [54] S . Nonaka and Y. Neba, Current regulated PWM-CSI induction motor drive system without a speed sensor, IEEE Transactions on Industry Applications, vol. 30, no. 1, pp. 116-125, 1994.

Control.

50

IEEE Control Systems

[55] P.D. Oliver, Feedback linearization of dc motors, IEEE Transactions on Industrial Electronics, vol. 38, no. 6, pp. 498-501, 1991. [56] R. Ortega and G. Espinosa, A controller design methodology for system with physical structures:Application to induction motors, Proceedf the 30th IEEE Conference on Decision and Control, Brighton, ings o England, pp. 2345-2349, 1991. [57] R. Ortega and G. Espinosa, Torque regulation of induction motors, Automatica, vol. 29, no. 3, pp. 621-633, 1993. [58] R. Ortega, C. Canudas, and S.I. Seleme, Nonlinear control of induction motors: Torque tracking with unknown load disturbance,IEEE Transactions on Automatic Control, vol. 38, no. 11, pp. 1675-1680, 1993. [59] R.H. Park, Two reaction theory of synchronousmachines: Part 1,AIEE Transactions, vol. 48, pp. 716-730, 1929. [60] R.H. Park, Two reaction theory of synchronous machines: Part 11, AIEE Transactions, vol. 52, pp. 352-354, 1933. [61] R.B. Sepe and J.H. Lang, Real-time adaptive control of the permanentmagnet synchronous motor, IEEE Transactions on Industry Applications, vol. 27, no. 4, pp. 706-714, 1991. [62] R.B. Sepe and J.H. Lang, Real-time observer-based (adaptive)control of a permanent-magnet synchronous motor without mechanical sensors, IEEE Transactions on Industry Applications, vol. 28, no. 6, pp. 1345-1352, 1992. [63] P.M. Sharkey and J. OReilly, Exact design manifold control of a class of nonlinear singularly perturbed system, IEEE Transactions on Automatic Control, vol. 32, no. 10, pp. 933-935, 1987. [64] K.R. Shouse, Reduced-order block techniques for singularly perturbed systems with application to permanent-magnet synchronousmotors, Ph.D. Thesis, Georgia Institute of Technology, Atlanta, GA, August 1993. [65] K.R. Shouse and D.G. Taylor, A digital self-tuning tracking controller for permanent-magnet synchronous motors, Proceedings of the 32nd IEEE pp. 3397-3402,1993. Conferenceon Decision and Control,San Antonio, TX, [66] K.R. Shouse and D.G. Taylor, Sensorless velocity control of permanent-magnet synchronous motors, to appear in Proceedings of the 33rd IEEE Conference on Decision and Control, Orlando, FL, 1994. [67] J.J.E. Slotine and W. Li, Applied Nonlinear Control. EngIewood Cliffs, NJ Prentice Hall, 1991. [68] R.C. Speagle and D.M. Dawson, Robust tracking control of a permanent magnet stepper motor driving a mechanical load, Proceedings of the Southeastem Symposium on System Theory, Tuscaloosa, AL, pp. 43-47, 1993. [69] R.C. Speagle and D.M. Dawson, Adaptive tracking control of a permanent magnet steppermotor driving a mechanical load, Proceedingsof the IEEE Southeastcon 93, Charlotte,NC, 1993 (no page numbers). [70] J. Stephan, M. Bodson, and J. Chiasson, Real-time estimation of the parameters and fluxes of induction motors, IEEE Transactions on Industry Applications, vol. 30, no. 3, pp. 746-759, 1994.

[71] H. Tajima and Y. Hori, Speed sensorless field-orientation control of the induction machine, IEEE Transactions on Industry Applications, vol. 29, no. 1, pp. 175-180, 1993. [72] D.G. Taylor, Pulse-width modulated control of electromechanical systems, IEEE Transactions on Automatic Control, vol. 37, no. 4, pp. 524-528,1992. [73] G.C. Verghese, J.H. Lang, and L.F. Casey, Analysis of instability in electrical machines, IEEE Transactions on Industry Applications, vol. 22, no. 5, pp. 852-864, 1986. [74] G.C. Verghese and S.R. Sanders, Observers for flux estimation in induction machines, IEEE Transactions on Industrial Electronics, vol. 35, no. 1, pp. 85-94, 1988. [75] R.S. Wallace and D.G. Taylor, Low-torque-ripple switched reluctance motors for direct-drive robotics, IEEE Transactions on Robotics and Automation, vol. 7, no. 6, pp. 733-742, 1991. [76] R.S. Wallace and D.G. Taylor, A balanced commutator for switched reluctance motors to reduce toque ripple, IEEE Transactions on Power Electronics,vol. 7, no. 4, pp. 617-626, 1992. [77] X. Xu and D.W. Novomy, Selection of the flux reference for induction machine drives in the field weakening region, IEEE Transactions on Industry Applications, vol. 28, no. 6, pp. 1353-1358, 1992. [78] G. Yang and T.H. Chin, Adaptive speed identification scheme for a vector-controlled speed sensorless inverter-induction motor drive, IEEE Transactions on Industry Applications, vol. 29, no. 4, pp. 820-825, 1993. [79] D.C. Youla and J.J. Bongiomo, A Floquet theory of the general linear rotating machine, IEEE Transactions on Circuits and Systems, vol. 27, pp. 15-19,1980, [80] M. Zribi and J. Chiasson, Position control of a PM stepper motor by exact linearization, IEEE Transactions on Automatic Control, vol. 36, no. 5, pp. 620-625,1991.

David G. Ikylorwas born in Oak Ridge, Tenn., on Aug. 7, 1961. He received the B.S. degree from the University of Tennessee, Knoxville, in 1983 and the M.S. and Ph.D. degrees from the University of Illinois, UrbanaChampaign, in 1985 and 1988, respectively, all in electrical engineering. During the summers from 1981 to 1983, he was employed by Oak Ridge National Laboratory, Oak Ridge, Tenn., and IBM Corp., Charlotte, N.C. From 1983 to 1988, he held various fellowships and assistantships at the University of Illinois, Urbana-Champaign. Since 1988, he has been with the School of Electrical and Computer Engineering, Georgia Institute of Technology, Atlanta, where he is now an associate professor. His research interests include nonlinear control theory and its applications to electromechanical systems.

I
I I

December 1994

51

1 I

Anda mungkin juga menyukai