Vous êtes sur la page 1sur 11

352

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 53, NO. 2, APRIL 2006

Review of Position-Sensorless Operation of Brushless Permanent-Magnet Machines


Paul P. Acarnley and John F. Watson, Senior Member, IEEE
AbstractThe operation of a brushless permanent-magnet machine requires rotor-position information, which is used to control the frequency and phase angle of the machines winding currents. Sensorless techniques for estimating rotor position from measurements of voltage and current have been the subject of intensive research. This paper reviews the state of the art in these sensorless techniques, which are broadly classied into three types: motional electromotive force, inductance, and ux linkage. Index TermsBrushless machines, estimation, permanentmagnet (PM) machines, position measurement, reviews.

I. I NTRODUCTION HE CURRENT ow in the windings of a brushless permanent-magnet (PM) machine must be synchronized to the instantaneous position of the rotor, and therefore, the current controller must receive information about the position of the machines rotor. An auxiliary device (e.g., an optical encoder or resolver) may be used to measure rotor position, but there has been much interest in sensorless schemes, in which position information is derived by on-line analysis of the voltages and currents in the machine windings. Fig. 1 shows a schematic sensorless scheme with a position estimation functional block receiving measurements of machine voltages and currents and supplying rotor-position data to the current and commutation controller. The sensorless approach has several advantages: 1) only electrical connections to the machine are the main phase windings, so installation costs are minimized; 2) position-sensing function can be located with the other control electronics: it does not need to be sited adjacent to the machine, and therefore does not inhibit the operating temperature range; 3) absence of connecting leads prevents corruption of position data by electromagnetic interference; and 4) cost of a position-encoding device is avoided. The idea of position-sensorless operation of brushless machines was rst advanced by Frus and Kuo [1] using a technique known as waveform detection for deducing rotor position in voltage-fed variable-reluctance stepping motors by analysis of current waveforms. Since that time, a number of titles have been used to describe this area of technology. Following close on waveform detection came the term indirect position sensing, which was justied by the observation that position
Manuscript received June 16, 2004; revised October 12, 2004. Abstract published on the Internet January 25, 2006. P. P. Acarnley is with the School of Electrical, Electronic and Computer Engineering, University of Newcastle, Newcastle upon Tyne, NE1 7RU, U.K. (e-mail: paul@reseeds.com). J. F. Watson is with the School of Engineering, Robert Gordon University, Aberdeen, AB10 1FR, U.K. Digital Object Identier 10.1109/TIE.2006.870868

Fig. 1. Position estimation from measurements of voltage and current in the machine supply.

sensing came indirectly from voltage and current waveforms. Other authors have used the term direct position sensing, because rotor position was obtained directly from the machine and not from a separate encoder. Even the generic term sensorless could be considered misleading: the techniques are position sensorless, but usually require sensing of current and sometimes voltage. This paper reviews the extensive research on sensorless operation of brushless PM machines, emphasizing particularly recent journal publications. Section II describes the characteristics of PM machines and their specic position-information requirements. Sensorless techniques may be broadly categorized as: motional electromotive force (EMF), inductance or uxlinkage sensing, each of which is described in Sections IIIV. II. C HARACTERISTICS AND P OSITION -S ENSING R EQUIREMENTS OF B RUSHLESS PM M ACHINES The purpose of this section is to outline the basic characteristics of brushless PM machines with particular reference to the requirements for rotor-position information. A. Categories of Brushless PM Machine Brushless PM machine drives can be divided into two subcategories [2]. The rst uses continuous rotor-position feedback to supply the motor with sinusoidal voltages and currents by pulsewidth modulation of the dc supply voltage. The ideal motional EMF is sinusoidal, so that the interaction with sinusoidal currents produces constant torque with very low ripple. This type of drive system is called a PM ac drive, brushless ac drive, sinusoidal-fed PM drive, or sinusoidal brushless dc drive. The second category of PM motor drive is referred to as the brushless dc drive, trapezoidal brushless dc drive, or rectangular-fed drive. In the three-phase form, rectangular current blocks of duration 120 electrical are supplied to the machine, in which the ideal motional EMF is trapezoidal, with

0278-0046/$20.00 2006 IEEE

ACARNLEY AND WATSON: POSITION-SENSORLESS OPERATION OF BRUSHLESS PERMANENT-MAGNET MACHINES

353

Fig. 2.

Equivalent circuit for one phase of a brushless PM machine.

the constant part of the waveform timed to coincide with the intervals of constant phase current. For this type of machine, rotor-position information is needed only at the commutation points, e.g., every 60 electrical in the three-phase machine. Both the trapezoidal and sinusoidal machines can be represented by the same equivalent circuit for each phase winding (Fig. 2), in which the source voltage v supplies current i to the phase circuit consisting of series-connected resistance R, inductance L, and motional EMF e. The motional EMF is caused by the movement of the PM rotor and is therefore dependent on rotor position, as well as being proportional to rotor velocity. Power supplied to the EMF e by the current i is converted into mechanical output power when the machine is acting as a motor. The rotor-position dependence of the inductance and motional EMF impacts on the machine voltage and current waveforms. This inter-relationship is utilized in sensorless schemes with the voltage and current waveforms being analyzed to extract the EMF or inductance (or a combination of the two), from which the rotor position is deduced. The torque output of a brushless PM motor is constant over a speed range limited by the power electronic converters ability to maintain the phase currents at the demanded levels. Fast and accurate control of phase-winding current is possible only if the available supply voltage from the dc link is substantially greater than the motional EMF, so that a surplus voltage is available to force current changes. The speed at which the surplus voltage is no longer adequate is referred to as the base speed. The machine can run above base speed in the eld-weakening mode, in which a component of the armature current produces a eld opposing the magnet eld and reduces the effective motional EMF. Field weakening is accomplished in both sinusoidal and trapezoidal machines by increasing the phase angle by which the current leads the motional EMF [3]. This requirement has important implications for position sensing: If the machines torqueproducing capability is to be fully utilized over a broad range of speeds in the eld-weakening region, then high-resolution position sensing is needed, even for trapezoidal machines. B. PM Machine Congurations Position variation of inductance and motional EMF in the PM machine depends on the magnetic structure. Brushless PM machines are characterized by having a eld produced by the permanent magnet on the rotor and the armature winding on

the stator. For the conventional inner rotor radial-eld machine, there are four possible rotor structures, as shown in Fig. 3. In the surface-mounted magnet arrangement [Fig. 3(a)], using modern rare-earth materials with magnetic relative permeability close to unity, the effective air gap is equal to the sum of the physical airgap between rotor and stator plus the magnet depth. Consequently, current owing in armature conductors produces only a small magnetic-ux component, and therefore, the inductance of the phase winding is small. Furthermore, if the entire rotor surface is covered by a permanent magnet, there is negligible variation in winding inductance with rotor position. The inset magnet conguration [Fig. 3(b)] is often preferred for trapezoidal machines, because the magnet pole arc can be adjusted to assist in shaping the motional-EMF waveform. The presence of soft-magnetic material at the physical airgap in the regions between the magnet poles causes a substantial variation in winding inductance, with maximum inductance occurring at rotor positions where the magnet pole arcs are misaligned from the winding axis. The two other congurations [Fig. 3(c) and (d)] have magnets buried in the rotor body. For the interior magnet structure [Fig. 3(c)], the direction of magnetization is radial. This structure is preferred for sinusoidal PM machines, because it is easier to achieve the necessary sinusoidal variation of ux density around the airgap periphery. The high-permeability magnetic material adjacent to the airgap leads to higher values of machine inductances than those that occur in the rst two congurations. Finally, the ux-concentrating type [Fig. 3(d)] has magnets located with their axes in the circumferential direction, so that the ux over a rotor pole arc is contributed by two separate magnets. This conguration also exhibits signicant saliency effects, causing a substantial variation of winding inductance with rotor position.

III. P OSITION S ENSING U SING M OTIONAL EMF A. Principles In brushless PM machines, movement of the magnets relative to the armature windings causes a motional EMF to be induced. Since the instantaneous magnitude of the EMF is a function of rotor position relative to the winding, information about position is contained in the EMF waveform. In practice, however, it is difcult to extract the information about EMF, because the machine windings are carrying rapidly changing currents and experience substantial induced voltages from phase switching. A further obstacle is that the motional EMF is proportional to rotor speed. When the machine is to operate from standstill, position sensing is possible only when a threshold speed is attained, so it is usual practice to make the initial acceleration under open-loop control using a ramped frequency signal [4], [5], the parameters of which must be chosen to match drive and load parameters. Acceleration from rest at an arbitrary start position for the rotor is particularly problematic, as it is possible for the motor to run initially in the reverse direction, and so, some schemes involve auxiliary sensors or high-frequency excitation tests to establish the rotor position and dene the appropriate initial winding excitation pattern.

354

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 53, NO. 2, APRIL 2006

Fig. 3. Rotor congurations of four-pole radial-eld machines. (a) Surface mounted. (b) Inset. (c) Interior. (d) Flux concentrating.

The simplest form of this sensing arrangement, reported by Iizuka et al. [6], can be understood by referring to the excitation timing diagram for a trapezoidal machine in the upper part of Fig. 4. The zero-crossing values of the motional EMF in each phase are an attractive feature to use for sensing, because they are independent of speed and because they occur at rotor positions where the phase winding is not excited. However, as shown in Fig. 4, the EMF zero crossings do not correspond to those rotor positions where commutation between phases should take place. Therefore, the signals must be phase shifted by 90 electrical before they can be used for commutation. For example, in Fig. 4, the shifted positive-going zero crossing of the phase-a motional EMF is used to trigger commutation of negative current from phase b to c. The simple motional-EMF sensing scheme described above has a number of limitations that prevent its general application. 1) In common with all motional-EMF schemes, sensing is not possible at low speeds. There are two restrictions on low-speed operation. First is the absence of motional EMF at zero speed, which can be addressed by accelerating the motor to a suitable speed with a preset sequence of excitation. A second factor is the requirement to phase shift the zero-crossing signals by 90 electrical. In the original work [6], the phase shifting was implemented using three separate RC networks that produce the necessary phase shift only when the working frequency is sufciently high. The heavy ltering inherent in the signal processing limits the dynamic performance of the position-sensing scheme. 2) It is assumed that there is a very rapid decay of current when a phase is switched off, so the voltage appearing across the terminals of the unexcited phase is equal to the motional EMF. This assumption may not be true at speeds approaching the base speed, or in the eld-weakening region. Therefore, there is an upper limit on the useful speed range obtainable with this form of motional-EMF sensing. 3) The motional EMF is measured across the terminals of each of the three machine phases in turn. For a starconnected machine, it is necessary to have a connection to the machines star point, and therefore there are four, rather than the conventional three, machine connections. Despite these restrictions, the method has been successfully applied in special-purpose low-cost applications for fans and pumps with unidirectional operation. For example,

Fig. 4. Derivation of current commutation signals from motional EMF (e) in a trapezoidal PM motor.

Iizuka et al. [6] described a 1.2-kW four-pole brushless dc motor for an air-conditioning unit with motional-EMF position sensing over a speed range from 1950 to 5700 r/min. An

ACARNLEY AND WATSON: POSITION-SENSORLESS OPERATION OF BRUSHLESS PERMANENT-MAGNET MACHINES

355

Fig. 5. Voltage reference points and current variables in a three-phase starconnected machine.

application of motional-EMF sensing in a brushless drive for an automotive fuel pump has been presented by Shao et al. [7], who overcame the starting problem with an experimentally determined open-loop excitation ramp. Toliyat et al. [8] described position sensing in a surface-mount magnet machine with insignicant saliency. This development involved a tapped machine winding that facilitated cancellation of both the thirdharmonic component of motional EMF and the resistive voltage drop, so there were no errors caused by changes in the winding resistance due to heating. Sensorless operation at excitation frequencies as low as 2 Hz was described. A digital phaselocked loop has been used by Amano et al. [9] to tackle the phase-shifting problem, though the dynamic range of sensorless operation was limited by the low-pass characteristics of the loop lter. A similar limitation has been experienced by others tackling the same problem [10][12]. For interior magnet machines, armature reaction may cause distortion of the airgap ux distribution and consequent errors in rotorposition detection. Shen and Tseng [13] have analyzed the error mechanism and developed a technique for armature-reaction compensation. Most of the work on motional-EMF sensorless operation has been implemented in a laboratory environment using either hard-wired analog/digital circuits or digital signal processors (DSPs). However, Cheng and Tzou [5] have designed and tested a mixed-mode integrated circuit, in a standard 0.35-m single-poly four-metal CMOS process, to perform all aspects of motional-EMF detection. In addition, several commercial integrated circuits are available, e.g., those in [14]. B. Sensing From the Third Harmonic of Motional EMF Moriera [15] introduced an improved method for position sensing using motional EMF, which utilizes the third-harmonic component in the EMF waveform of a trapezoidal PM machine, and thus reduces the phase-shifting problem outlined above. Fig. 5 shows the three windings of a star-connected machine with star point s. An additional star connection of three identical resistors is connected between the phase ends a, b, c, and a separate star point n. Assuming that the resistances I and inductances (L) of the three machine windings are identical and that the phase motional EMFs are as shown in Fig. 6, it can be shown that the voltage between the two star points n and s is equal to the mean of the three phase EMFs. Fig. 6 shows the voltage vns and its relationship to the rotor positions for switching between phases. The vns waveform has a frequency three times that of the fundamental component

Fig. 6. Derivation of current commutation signals from third harmonic of motional EMF (e) in a trapezoidal PM motor.

of any of the phase motional EMFs, and therefore, it is referred to as the third harmonic of motional EMF, though it also contains harmonics of higher orders. The waveform is shifted through a rotor position of 30 by integration. The zero crossings of the integrated waveform correspond to the rotor positions at which excitation must be switched between the phases, and therefore, the zero crossings are suitable excitation switching signals.

356

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 53, NO. 2, APRIL 2006

Fig. 7. Principles of a closed-loop observer.

During each cycle of excitation, the vns waveform passes through three cycles. Therefore, the excitation trigger signals must be synchronized to the appropriate phase excitation change, depending on the required directions of torque and speed. The trigger signals are synchronized once per excitation cycle by identifying a suitable reference position, such as the positive-going zero crossing of motional EMF in phase a. In comparison to the basic method of position sensing using motional EMF (Section III-A), the third-harmonic method has the following advantages. 1) There is a reduced ltering requirement, because the integration (low-pass ltering) function is performed on a signal, which has a frequency three times that of the fundamental signal. The lighter ltering improves the dynamic performance. 2) Operation at higher speeds is possible in principle, because the voltage vns can be recovered even if current continues to ow in the third (unexcited) phase. Shen et al. [16] have successfully applied third harmonic of back EMF sensing in the high-speed ux-weakening region with current owing continuously in all three phases of the machine. An important limitation on the third-harmonic approach is the initial assumption in the analysis that the inductance is equal in all three phases. This assumption is usually valid for surface-mounted magnet machines, but is not correct for rotor congurations that cause the machine to exhibit signicant saliency. In the latter case, errors in position estimation arise due to rapidly changing phase currents and extra low-pass ltering may be needed. C. Observer-Based Methods Observer principles (Fig. 7) have been applied to sensorless operation of PM machines. A machine and power converter are supplied with one or more inputs (e.g., voltages) and produce several measured outputs (e.g., currents). A mathematical model of the converter/machine combination is supplied with the same inputs and produces estimates of the outputs. These estimated outputs are compared with the measured outputs to yield an estimation error, which is fed back to the model to assist in correcting the estimates. If the estimation error is small, the model replicates the behavior of the real converter

and machine. All of the states in the mathematical model are accessible, so estimates of all physical quantities are available, including those which are difcult or expensive to measure (e.g., rotor position or phase ux linkage). Closed-loop observers have been used to address position sensing in PM machines [17][25]. Solsana et al. [17] described the application of the principle to a sinusoidal PM machine and presented simulation results showing how the machine can start successfully even when there were initial errors in estimated speed and rotor position. A later paper by the same authors [18] discussed the impact of modeling errors, such as space harmonics of the magnet eld, on the estimation of rotor speed. Two alternative models (a voltage model and a current model) for a PM machine with signicant rotor saliency have been used for position estimation by Matsui and co-worker [19], [20]. However, the starting problem remained, and for consistent starting, a separate technique was needed [21], often using the inductance-variation approach described in Section IV. In applications where erratic initial motion can be tolerated, modied control laws have been proposed [22] for robust sensorless control at low speeds. A model-based approach to EMF sensing has been described by Cho et al. [23] for a surface-mounted magnet machine for a direct drive in a washing machine. In this application, the machine was subjected to large temperature variations, which impacted on both the stator resistance and the remnant ux density of the ferrite magnets. Therefore, the machines temperature was estimated, via the stator resistance, at those parts of the wash cycle where the rotor speed was zero, and appropriate corrections were applied to the model parameters. An alternative approach to temperature-dependent parameter tracking [24] involved the injection of a current perturbation signal when the machine was operating at steadystate speed. IV. P OSITION S ENSING U SING I NDUCTANCE V ARIATION An alternative method of position sensing involves monitoring rates of change of winding current. Since the rate of current change depends on the inductance of the winding, and this inductance is a function of rotor position and winding current, then rotor position can be deduced from winding current and its rate of change. Such a scheme has the important advantage that it is useful even at zero speed where there is no motional EMF. Sensing of rotor position by inductance variation in the brushless PM machine is complicated because: 1) in a machine with surface-mounted magnets, there is no inherent saliency, so any variation of winding inductance with rotor position arises only from magnetic saturation; 2) the rate of change of current in the PM machine is dominated by the motional EMF; 3) the variation of incremental inductance with rotor position undergoes two cycles per single electrical cycle of the PM machine, causing an ambiguity in sensed position. The nal point is illustrated in Fig. 8, which shows the variation of the various components of ux linkage over two cycles of a two-pole PM machine. The minimum value of incremental inductance occurs at rotor positions of both 0 and 180 , but the magnet ux linkage is maximum positive at

ACARNLEY AND WATSON: POSITION-SENSORLESS OPERATION OF BRUSHLESS PERMANENT-MAGNET MACHINES

357

Fig. 9. Current pulse amplitude as a function of rotor position in a machine with magnetic saturation.

Fig. 8. Flux linkages and incremental inductance as a function of rotor position in a PM machine with saliency.

the position of 0 and maximum negative at the 180 position. Despite these obvious difculties, there have been numerous attempts to use inductance variation to detect rotor position in PM machines. The rst application of inductance methods addresses the problem of starting, including identication of the rotor position before full excitation is applied to the machine. Initial position identication is particularly important in applications such as traction, where any reverse motion caused by incorrect excitation is unacceptable. Exploratory voltage signals have been applied to the phase windings of a salient PM machine that was stationary [20], [26][29]. The resulting current amplitudes depended on the incremental inductance, and therefore the rotor position. However, the sensed position spanned a range of 180 electrical and there remained the problem of resolving the orientation of the rotor over the full 360 electrical range. The universally adopted solution to the issue of rotor orientation ambiguity is to consider the effect of magnetic saturation on the incremental inductance. The principles of this method can be understood by referring to the ux-linkage characteristics in Fig. 8. Suppose the rotor position has been determined as either 0 or 180 , by observing that the incremental inductance of a phase is at its minimum value. Now, consider the effect of a positive pulse of current. If the rotor is aligned at the 0

position, the effect of the pulse is to increase the total positive ux linked with the phase, but if the rotor is at the 180 position, the current reduces the total negative ux linkage. Therefore, there is a difference between the ux amplitudes for the two alternative rotor positions and, consequently, a difference in the level of magnetic saturation. If magnetic saturation increases, the incremental inductance is lower and so the amplitude of the current pulse is larger at one of the two possible rotor positions, as illustrated in Fig. 9. Magnetic saturation has a small but signicant inuence on incremental inductance even in a surface magnet machine with no inherent saliency effect. The study by Nakashima et al. [30] has been successful in utilizing saturation effects to estimate initial rotor position in such machines, albeit with reduced accuracy: a maximum error of 18 electrical was reported. The initial starting position of a salient PM motor can be investigated by high-frequency injection methods that effectively detect the position-dependent incremental inductance [31], [32], [34][36]. For example, Noguchi et al. [31] evaluated the winding impedance with the pulsewidth-modulated voltage controller producing a low-amplitude output at a frequency of 50 Hz. Ambiguity in the magnet direction was resolved by adjusting the parameters of the closed-loop current controller so that an oscillatory current response was obtained for the lowest values of incremental inductance, corresponding to maximum saturation. Aihara et al. [32] used a higher frequency (500 Hz) for signal injection and discriminated between magnet polarities by means of rotor magnetic-saturation effects. The initial rotor position can be detected using voltage pulses, as described by Lai et al. [33]. In this paper, the dc-link voltage was applied to various combinations of the three starconnected phases using the six-device bridge. The current was allowed to build up to a steady-state level in all phases, but when the voltage was removed, current decayed faster in the phase with lowest inductance. There was no need for phase-current sensors, because the current decay was evaluated by means of the line voltages, which underwent a sudden change in level when the corresponding phase currents fell to zero. Several methods of position sensing from inductance variation have been developed for continuous rotation of salient machines. Kulkarni and Ehsani [37] proposed a method for calculating the effective phase inductance from the behavior of the hysteresis current controller for an excited phase. Ambiguity in the sensed position was avoided by starting the machine from a known location and continuously tracking changes in inductance with the assumption that the machine was always

358

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 53, NO. 2, APRIL 2006

rotating in one direction. Corley and Lorenz [38] utilized voltage injection at a carrier frequency of 2 kHz. The corresponding frequency component of current was modulated by the rotorposition variation of the phase inductance. Information about the rotor position was extracted by making a comparison with a set of signals at the same carrier frequency and modulated by an estimated rotor position, which was derived from a simple motion estimator. The result of the comparison was a signal modulated by the error between the actual and estimated rotor positions. The position-error information was decoded and used to correct the motion estimators output, which therefore tracked the actual rotor position over a wide speed range, including zero speed. A similar approach has been reported by Noguchi and Kohno [39], who used a 4-kHz carrier frequency, while Shinnaka [40] developed a new ltering algorithm, based on a phase-locked loop, which exhibited good low-speed performance. Sensorless operation from inductance variation has been achieved using injected voltage sequences during breaks in the normal voltage modulation [41], and from the pulsewidthmodulation frequency components at switching frequencies up to 5.5 kHz [42], [43]. There is a tradeoff in the choice of modulation frequency: a low frequency results in larger more easily detected current amplitudes, but can cause audible noise from the motor, whereas a high modulation frequency avoids audible noise, but the current amplitudes are much reduced. Choeisai et al. [44] were able to work with a modulation frequency of 20 kHz by using a Walsh detector to assist in measuring the high-frequency current amplitudes. Hartas et al. [45] developed a novel six-phase machine and implemented sensorless operation from the position-dependent behavior of the hysteresis current controller. A detailed investigation of high-frequency signal injection, using experimental and niteelement analysis techniques, by Jang et al. [46] has highlighted the need for careful choice of signal amplitude and frequency. A further option for improving position sensing is to modify the machine rotor. For example, Nondahl et al. [47] added a short-circuited rotor winding to a surface magnet machine with no inherent saliency. The winding current reacted with time-varying elds acting in the direction of the main magnet eld, and therefore enhanced the position dependence of the winding inductance. A similar effect has been obtained by coating rotor poles of like polarity with a nonmagnetic conducting material [48]. Matsuse et al. [49] used a new closedslot stator conguration to assist in rotor-position estimation. Magnetic bridges across the stator slots included a region of constricted cross section. These regions conducted a component of winding leakage ux, but become saturated by the rotor magnet ux, which therefore inuenced the phase-winding inductances. V. P OSITION S ENSING B ASED ON F LUX -L INKAGE V ARIATION A. Principles Position sensing via ux-linkage variation has been known for many years, but its successful implementation has become

possible only in the last decade with the emergence of devices with sufcient real-time processing power. The fundamental idea of ux-linkage position sensing is deceptively simple. The phase-voltage equation can be written as v =Ri+ d dt (1)

where v phase terminal voltage; i phase current; R phase resistance; phase ux linkages; and the phase ux linkages are a function of current and rotor position. Equation (1) can be rearranged as = {v R i} dt. (2)

Therefore, it appears that by subtracting the resistive voltage drop from the phase voltage and integrating, then a continuous estimate of phase ux linkage can be produced. In most electrical machines, it is not practicable to measure the phase terminal voltages directly, because of isolation issues, and instead, the applied phase voltage is estimated from knowledge of the solid-state converter dc supply voltage and demands being input to the voltage controller. An important source of error here is the conversion of voltage demands into converter device switching signals, which must include some dead time between switching off one device in an inverter phase leg and switching on the other device in the same leg. The effect of dead time is to introduce an error between demanded and actual values of phase voltage, with the error being greatest at output voltages near zero. Several authors have highlighted this voltage measurement error and have developed techniques for compensating dead-time errors [50], [51]. The open-loop integration in (2) is prone to errors caused by drift: Small offset signals in the measurements are summed over time, causing the integrator output to saturate. Integrator drift can be reduced if the pure integrator is replaced by a low-pass lter or an alternative integrator structure [53], but this modication inhibits the ux estimators low-speed operating range. Instead, the general trend of recent research on this aspect of sensorless operation has been to concentrate on closed-loop ux-linkage estimation as part of the position-sensing process. This section examines position-sensorless operation using ux linkages and draws a distinction between those estimators that include a mechanical system model, and therefore require knowledge of mechanical parameters, and estimators that are independent of a mechanical model, and are therefore more suitable for operation with variable load conditions. Position estimation using ux linkages may be viewed as an amalgamation of the sensing methods using motional EMF (Section III) and inductance (Section IV): Several authors have married the concepts by introducing the idea of an extended EMF (EEMF) estimator, which includes motional EMF and inductive terms [54][56]. However, it should be emphasized that ux-linkage estimation does not access any more position information than is available from the combination of motional

ACARNLEY AND WATSON: POSITION-SENSORLESS OPERATION OF BRUSHLESS PERMANENT-MAGNET MACHINES

359

Fig. 10. Closed-loop observer for position estimation using ux linkages ( = estimated state).

EMF and position sensing, so there remain machine types for which position-sensorless operation is extremely difcult under certain operating conditions. B. Flux Estimation With a Mechanical Model In the brushless PM machine, the ux linkages with each phase arise from the permanent magnet itself and the currents owing in the machine windings. The magnet ux linkages are a function of rotor position with the nature of the function depending on the envisaged operating mode of the machine. In a brushless dc machine, the magnet ux linkages are a trapezoidal function of position, but in the sinusoidal machine, a sinusoidal variation of magnet ux linkage with position is needed. As noted in Section IV, the phase-inductance variation, and hence the variation of ux linkages from the phase current, depends on the machine construction. Flux-linkage approaches to position estimation in all types of machines frequently make use of the closed-loop observer principles shown in Fig. 10. Voltage and current supplied to a machine are input to an estimator, which rst calculates values of ux linked with each phase. Flux linkage and current can be used to calculate torque, and the torque data are input to a mechanical model 1 d = (T B TL ), dt J d = dt (3)

been developed by Terzic and Jadric [51] and implemented in a DSP. Stator-resistance estimation was also included in the algorithm to counter the effects of ux-linkage estimation errors caused by an incorrect value of resistance as the motor temperature rose during continuous operation. Sensorless operation at speeds down to 50 r/min was recorded for a machine with negligible saliency. An obstacle to applying the extended Kalman lter algorithm to rotor-position estimation is the need to set appropriate values for the covariance matrix parameters, which reect the uncertainties in modeling and measurements. The parameter values are often chosen by trial and error, but Bolognani et al. [52] have developed selection guidelines. C. Flux Estimation Without a Mechanical Model Closed-loop observers that include a model of the mechanical dynamics are prone to errors caused by incorrect specication or variability of the mechanical parameters. Therefore, a number of methods for avoiding a mechanical model have been introduced. Position estimation from ux-linkage variation was investigated by Wu and Slemon [59] for the case of a sinusoidal machine without signicant saliency. Analog integrators were used for ux-linkage calculation, because the inverter was hysteresis controlled, so it would have been difcult to synchronize data sampling to randomly occurring switching events. Integrator drift was counteracted by calculation of an offset signal, which ensured that the average ux linkage with each phase was zero when the machine operated at steady-state speed and torque. This approach was an effective low-cost solution, but did not allow the machine to self-start and was unable to deal with fast changes in speed or torque. The principles of ux linkage and position estimation with a closed-loop observer but without the need for a mechanical model are represented schematically in Fig. 11. Flux linkage is calculated by integration of the input voltage and current, and the stored ux linkage/position/current characteristic is used to estimate current and rotor position. The estimated current is fed back for use in the lookup block, as well as being compared to the measured currents to generate an estimation error. This closed-loop feature counteracts integrator drift.

(where: J = mechanical inertia, B = viscous friction constant, TL = load torque). Estimates of rotor speed and position are produced. To close the estimation loop, the estimated values of position and phase ux linkages are used to estimate the phase currents, which are compared with the measured currents to generate an estimation error. This error is fed back into the initial ux-linkage estimator, where it counteracts the tendency to drift caused by offsets in the measurements. Initial attempts at combined ux linkage and position observation for brushless PM machines by Jones and Lang [57] and Sepe and Lang [58] were frustrated by the limits on available real-time processing power at that time. The subsequent advances in DSP technology have allowed these observer principles to be implemented effectively. For example, an extended Kalman lter estimator for a brushless dc motor has

360

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 53, NO. 2, APRIL 2006

dc PM machines is there an established commercial interest, with the major manufacturers producing integrated circuits for sensorless operation using motional EMF. However, further reductions in the cost of processing power will improve the economic viability of many of the more sophisticated sensorless techniques described here. R EFERENCES
[1] J. R. Frus and B. C. Kuo, Closed-loop control of step motors using waveform detection, in Proc. Int. Conf. Stepping Motors and Systems, Leeds, U.K., 1976, pp. 7784. [2] P. Pillay and R. Krishnan, Application characteristics of permanent magnet synchronous and brushless DC motors for servo drives, IEEE Trans. Ind. Appl., vol. 27, no. 5, pp. 986996, Sep./Oct. 1991. [3] S. K. Sa, P. P. Acarnley, and A. G. Jack, Analysis and simulation of the high-speed torque performance of brushless DC motor drives, Proc. Inst. Elect. Eng.Electr. Power Appl., vol. 142, no. 3, pp. 191200, Mar. 1995. [4] H.-C. Chen and C.-M. Liaw, Current-mode control for sensorless BDCM drive with intelligent commutation tuning, IEEE Trans. Power Electron., vol. 17, no. 5, pp. 747756, Sep. 2002. [5] K.-Y. Cheng and Y.-Y. Tzou, Design of a sensorless commutation IC for BLDC motors, IEEE Trans. Power Electron., vol. 18, no. 6, pp. 1365 1375, Nov. 2003. [6] K. Iizuka, H. Uzuhashi, M. Kano, T. Endo, and K. Mohri, Microcomputer control for sensorless brushless motor, IEEE Trans. Ind. Appl., vol. IA-21, no. 4, pp. 595601, May/Jun. 1985. [7] J. Shao, D. Nolan, M. Teissier, and D. Swanson, A novel microcontrollerbased sensorless brushless DC (BLDC) motor drive for automotive fuel pumps, IEEE Trans. Ind. Appl., vol. 39, no. 6, pp. 17341740, Nov./Dec. 2003. [8] H. A. Toliyat, L. Hao, D. S. Shet, and T. A. Nondahl, Position-sensorless control of surface-mount permanent-magnet AC (PMAC) motors at low speeds, IEEE Trans. Ind. Electron., vol. 49, no. 1, pp. 157164, Feb. 2002. [9] Y. Amano, T. Tsuji, A. Takahashi, S. Ouchi, K. Hamatsu, and M. Iijima, A sensorless drive system for brushless DC motors using a digital phaselocked loop, Electr. Eng. Jpn., vol. 142, no. 1, pp. 5766, 2003. [10] G.-J. Su and J. W. McKeever, Low-cost sensorless control of brushless DC motors with improved speed range, IEEE Trans. Power Electron., vol. 19, no. 2, pp. 296302, Mar. 2004. [11] , Low-cost sensorless control of brushless DC motors with improved speed range, IEEE Trans. Power Electron., vol. 19, no. 3, pp. 878879, May 2004. [12] D.-H. Jung and I.-J. Ha, Low-cost sensorless control of brushless DC motors using a frequency-independent phase shifter, IEEE Trans. Power Electron., vol. 15, no. 4, pp. 744752, Jul. 2000. [13] J. X. Shen and K. J. Tseng, Analyses and compensation of rotor position detection error in sensorless pm brushless DC motor drives, IEEE Trans. Energy Convers., vol. 18, no. 1, pp. 8793, Mar. 2003. [14] J. P. M. Bahlmann, A full-wave motor drive IC based on the backemf sensing principle, IEEE Trans. Consum. Electron., vol. 35, no. 3, pp. 415420, Aug. 1989. [15] J. C. Moreira, Indirect sensing for rotor ux position of permanent magnet AC motors operating over a wide speed range, IEEE Trans. Ind. Appl., vol. 32, no. 6, pp. 13941401, Nov./Dec. 1996. [16] J. X. Shen, Z. Q. Zhu, and D. Howe, Sensorless ux-weakening control of permanent-magnet brushless machines using third harmonic back emf, IEEE Trans. Ind. Appl., vol. 40, no. 6, pp. 16291636, Nov./Dec. 2004. [17] J. Solsona, M. I. Valla, and C. Muravchik, A nonlinear reduced order observer for permanent magnet synchronous motors, IEEE Trans. Ind. Electron., vol. 43, no. 4, pp. 492497, Aug. 1996. [18] , On speed and rotor position estimation in permanent-magnet AC drives, IEEE Trans. Ind. Electron., vol. 47, no. 5, pp. 11761180, Oct. 2000. [19] N. Matsui and M. Shigyo, Brushless DC motor control without position and speed sensors, IEEE Trans. Ind. Appl., vol. 28, no. 1, pp. 120127, Jan./Feb. 1992. [20] N. Matsui, Sensorless pm brushless DC motor drives, IEEE Trans. Ind. Electron., vol. 43, no. 2, pp. 300308, Apr. 1996. [21] T. Takeshita, A. Usui, and N. Matsui, Sensorless salient-pole PM synchronous motor drive in all speed ranges, Electr. Eng. Jpn., vol. 135, no. 3, pp. 6473, 2001.

Fig. 11. Position estimation using ux linkages without a mechanical model ( = estimated state).

Batzel and Lee [60] used a Luenberger observer for ux linkage and position estimation in a sinusoidal PM machine, which was both slotless and had surface-mounted rotor magnets. The ux linkage/position/current characteristic was represented by a sinusoidal characteristic for the magnet ux linkages. In the absence of saliency, position estimation was not possible at the lowest speeds, so Hall sensors were used to provide an auxiliary position measurement with a resolution of 30 for operation below 100 r/min. A later work by the same authors [61] has involved a neural network for the real-time correction of position-estimation errors arising from uncertainty in the machine parameters. An alternative to a mechanical model in a ux linkage and position estimator is to obtain an initial estimate of rotor position, by extrapolation from previous position estimates, and then correct the position in response to measured voltage and current information. An example of this approach [62], [63] used a combined phase ux linkage and rotor-position estimator, in which integrator drift was counteracted by reconciling the estimates with measured values of current according to the stored ux-linkage/current/rotor-position characteristics of the machine. The same algorithm has been used successfully by Ostlund and Brokemper [26], who reported the need for an auxiliary starting method when using a surface-mounted magnet machine. Sensorless operation with ux-linkage estimation has been developed for the control of a PM motor with a matrix power converter [64]. The problems of ux integrator drift have been avoided [65] by considering the relationship between incremental changes in ux linkage and rotor position to establish a tracking algorithm requiring only integration of position increments. VI. C ONCLUSION Sensorless technology for PM machines has been researched extensively with a surge of recent interest being prompted by the availability of more powerful digital signal processing devices. Solutions have been found to the more difcult problems of sensorless operation, such as starting from rest and position sensing for surface-mounted magnet machines. Despite this extensive development, only in the very specic areas of lowcost [7], [23] or safety-critical applications [66] for brushless

ACARNLEY AND WATSON: POSITION-SENSORLESS OPERATION OF BRUSHLESS PERMANENT-MAGNET MACHINES

361

[22] B. Nahid-Mobarakeh, F. Meibody-Tabar, and F.-M. Sargos, Mechanical sensorless control of PMSM with online estimation of stator resistance, IEEE Trans. Ind. Appl., vol. 40, no. 2, pp. 457471, Mar./Apr. 2004. [23] K. Y. Cho, S. B. Yang, and C. H. Hong, Sensorless control of a PM synchronous motor for direct drive washer without rotor position sensors, Proc. Inst. Elect. Eng.Electr. Power Appl., vol. 151, no. 1, pp. 6169, Jan. 2004. [24] K.-W. Lee, D.-H. Jung, and I.-J. Ha, An online identication method for both stator resistance and back-emf coefcients of PMSMs without rotational transducers, IEEE Trans. Ind. Electron., vol. 51, no. 2, pp. 507510, Apr. 2004. [25] G. D. Andreescu, Adaptive observer for sensorless control of permanent magnet synchronous motor drives, Electr. Power Compon. Syst., vol. 30, no. 2, pp. 107119, Feb. 2002. [26] S. Ostlund and M. Brokemper, Sensorless rotor position detection from zero to rated speed for an integrated PM synchronous motor drive, IEEE Trans. Ind. Appl., vol. 32, no. 5, pp. 11581165, Sep./Oct. 1996. [27] J.-I. Ha, K. Ide, T. Sawa, and S.-K. Sul, Sensorless rotor position estimation of an interior permanent-magnet motor from initial states, IEEE Trans. Ind. Appl., vol. 39, no. 3, pp. 761767, May/Jun. 2003. [28] G. H. Jang, J. H. Park, and J. H. Chang, Position detection and startup algorithm of a rotor in a sensorless BLDC motor utilising inductance variation, Proc. Inst. Elect. Eng.Electr. Power Appl., vol. 149, no. 2, pp. 137142, Mar. 2002. [29] M. Tursini, R. Petrella, and F. Parasiliti, Initial rotor position estimation method for pm motors, IEEE Trans. Ind. Appl., vol. 39, no. 6, pp. 1630 1640, Nov./Dec. 2003. [30] S. Nakashima, Y. Inagaki, and I. Miki, Sensorless initial rotor position estimation of surface permanent-magnet synchronous motor, IEEE Trans. Ind. Appl., vol. 36, no. 6, pp. 15981603, Nov./Dec. 2000. [31] T. Noguchi, K. Yamada, S. Kondo, and I. Takahashi, Initial rotor position estimation method of sensorless PM synchronous motor with no sensitivity to armature resistance, IEEE Trans. Ind. Electron., vol. 45, no. 1, pp. 118125, Feb. 1998. [32] T. Aihara, A. Toba, T. Yanase, A. Mashimo, and K. Endo, Sensorless torque control of salient-pole synchronous motor at zero-speed operation, IEEE Trans. Power Electron., vol. 14, no. 1, pp. 202208, Jan. 1999. [33] Y.-S. Lai, F.-S. Shyu, and S. S. Tseng, New initial position detection technique for three-phase brushless DC motor without position and current sensors, IEEE Trans. Ind. Appl., vol. 39, no. 2, pp. 485491, Mar./Apr. 2003. [34] J.-H. Jang, S.-K. Sul, J.-I. Ha, K. Ide, and M. Sawamura, Sensorless drive of surface-mounted permanent-magnet motor by high-frequency signal injection based on magnetic saliency, IEEE Trans. Ind. Appl., vol. 39, no. 4, pp. 10311038, Jul./Aug. 2003. [35] M. E. Haque, L. Zhong, and M. F. Rahman, A sensorless initial rotor position estimation scheme for a direct torque controlled interior permanent magnet synchronous motor drive, IEEE Trans. Power Electron., vol. 8, no. 6, pp. 13761383, Nov. 2003. [36] H. Kim, K.-K. Huh, R. D. Lorenz, and T. M. Jahns, A novel method for initial rotor position estimation for IPM synchronous machine drives, IEEE Trans. Ind. Appl., vol. 40, no. 5, pp. 13691378, Sep./Oct. 2004. [37] A. B. Kulkarni and M. Ehsani, A novel position sensor elimination technique for the interior permanent-magnet synchronous motor drive, IEEE Trans. Ind. Appl., vol. 28, no. 1, pp. 144150, Jan./Feb. 1992. [38] M. J. Corley and R. D. Lorenz, Rotor position and velocity estimation for a salient-pole permanent magnet synchronous motor at standstill and high speeds, IEEE Trans. Ind. Appl., vol. 34, no. 4, pp. 784789, Jul./Aug. 1998. [39] T. Noguchi and S. Kohno, Mechanical-sensorless permanent-magnet motor drive using relative phase information of harmonic currents caused by frequency-modulated three-phase PWM carriers, IEEE Trans. Ind. Appl., vol. 39, no. 4, pp. 10851092, Jul./Aug. 2003. [40] S. Shinnaka, New mirror-phase vector control for sensorless drive of permanent-magnet synchronous motor with pole saliency, IEEE Trans. Ind. Appl., vol. 40, no. 2, pp. 599606, Mar./Apr. 2004. [41] E. Robeischl and M. Schroedl, Optimized INFORM measurement sequence for sensorless pm synchronous motor drives with respect to minimum current distortion, IEEE Trans. Ind. Appl., vol. 40, no. 2, pp. 591598, Mar./Apr. 2004. [42] V. Petrovic, A. M. Stankovic, and V. Blasko, Position estimation in salient PM synchronous motors based on PWM excitation transients, IEEE Trans. Ind. Appl., vol. 39, no. 3, pp. 835843, May/Jun. 2003. [43] S. Ogasawara, T. Matsuzawa, and H. Akagi, A position-sensorless IPM motor drive system using a position estimation based on magnetic saliency, Electr. Eng. Jpn., vol. 131, no. 2, pp. 6879, 1999.

[44] K. Choeisai, N. Kobayashi, and S. Kondo, Walsh function-based position sensorless control for interior permanent-magnet motor drives using ripple-current of high-frequency triangular-wave-carrier PWM inverter, Electr. Eng. Jpn., vol. 145, no. 3, pp. 8088, 2003. [45] P. Hartas, A. Shahin, K. I. Nuttall, and D. W. Shimmin, A novel permanent magnet DC motor and sensorless electrical drive system, Electr. Power Compon. Syst., vol. 29, no. 1, pp. 1528, Jan. 2001. [46] J.-H. Jang, J.-I. Ha, M. Ohto, K. Ide, and S.-K. Sul, Analysis of permanent-magnet machine for sensorless control based on highfrequency signal injection, IEEE Trans. Ind. Appl., vol. 40, no. 6, pp. 15951604, Nov./Dec. 2004. [47] T. A. Nondahl, G. Ray, P. B. Schmidt, and M. L. Gasperi, A permanentmagnet rotor containing an electrical winding to improve detection of rotor angular position, IEEE Trans. Ind. Appl., vol. 35, no. 4, pp. 819824, Jul./Aug. 1999. [48] M. Tomita, S. Doki, S. Okuma, and H. Yamaguchi, Sensorless rotor position estimation at standstill of cylindrical brushless DC motors using opened phase voltage change caused by eddy currents, Electr. Eng. Jpn., vol. 126, no. 1, pp. 5260, 1999. [49] K. Matsuse, T. Baba, I. Masukane, H. Ohta, and R. Tsuchimoto, Positionsensorless starting method and driving characteristics of closed-slot small permanent-magnet motor, IEEE Trans. Ind. Appl., vol. 39, no. 2, pp. 451456, Mar./Apr. 2003. [50] L. Harnefors and H. P. Nee, A general algorithm for speed and position estimation of AC motors, IEEE Trans. Ind. Electron., vol. 47, no. 1, pp. 7783, Feb. 2000. [51] B. Terzic and M. Jadric, Design and implementation of the extended Kalman lter for the speed and rotor position estimation of brushless DC motor, IEEE Trans. Ind. Electron., vol. 48, no. 6, pp. 10651073, Dec. 2001. [52] S. Bolognani, L. Tubiana, and M. Zigliotto, Extended Kalman lter tuning in sensorless PMSM drives, IEEE Trans. Ind. Appl., vol. 39, no. 6, pp. 17411747, Nov./Dec. 2003. [53] J. Hu and B. Wu, New integration algorithms for estimating motor ux over a wide speed range, IEEE Trans. Power Electron., vol. 13, no. 5, pp. 969977, Sep. 1998. [54] S. Ichikawa, Z. Chen, M. Tomita, S. Doki, and S. Okuma, Sensorless controls of salient-pole permanent magnet synchronous motors using extended electromotive force models, Electr. Eng. Jpn., vol. 146, no. 3, pp. 5564, 2004. [55] Z. Chen, M. Tomita, S. Doki, and S. Okuma, An extended electromotive force model for sensorless control of interior permanent-magnet synchronous motors, IEEE Trans. Ind. Electron., vol. 50, no. 2, pp. 288295, Apr. 2003. [56] S. Morimoto, K. Kawamoto, and Y. Takeda, Position and speed sensorless control for IPMSM based on estimation of position error, Electr. Eng. Jpn., vol. 144, no. 2, pp. 4352, 2003. [57] L. A. Jones and J. H. Lang, A state observer for the permanent magnet synchronous motor, IEEE Trans. Ind. Electron., vol. 36, no. 3, pp. 374382, Aug. 1989. [58] R. B. Sepe and J. H. Lang, Real-time observer-based (adaptive) control of a permanent-magnet synchronous motor without mechanical sensors, IEEE Trans. Ind. Appl., vol. 28, no. 6, pp. 13451352, Nov./Dec. 1992. [59] R. Wu and G. R. Slemon, A permanent magnet motor drive without a shaft sensor, IEEE Trans. Ind. Appl., vol. 27, no. 5, pp. 10051011, Sep./Oct. 1991. [60] T. D. Batzel and K. Y. Lee, Slotless permanent magnet synchronous motor operation without a high resolution rotor angle sensor, IEEE Trans. Energy Convers., vol. 15, no. 4, pp. 366371, Dec. 2000. [61] , An approach to sensorless operation of the permanent-magnet synchronous motor using diagonally recurrent neural networks, IEEE Trans. Energy Convers., vol. 18, no. 1, pp. 100106, Mar. 2003. [62] N. Ertugrul and P. P. Acarnley, New algorithm for sensorless operation of permanent magnet motors, IEEE Trans. Ind. Appl., vol. 30, no. 1, pp. 126133, Jan./Feb. 1994. [63] C. D. French and P. P. Acarnley, Control of permanent magnet motor drives using a new position estimation technique, IEEE Trans. Ind. Appl., vol. 32, no. 5, pp. 10891097, Sep./Oct. 1996. [64] T.-H. Liu, S.-H. Chen, and D.-F. Chen, Design and implementation of a matrix converter PMSM drive without a shaft sensor, IEEE Trans. Aerosp. Electron. Syst., vol. 39, no. 1, pp. 228243, Jan. 2003. [65] L. Ying and N. Ertugrul, A novel, robust DSP-based indirect rotor position estimation for permanent magnet AC motors without rotor saliency, IEEE Trans. Power Electron., vol. 18, no. 2, pp. 539546, Mar. 2003. [66] S. Green, D. J. Atkinson, A. G. Jack, B. C. Mecrow, and A. King, Sensorless operation of a fault tolerant PM drive, Proc. Inst. Elect. Eng.Electr. Power Appl., vol. 150, no. 2, pp. 117125, Mar. 2003.

362

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 53, NO. 2, APRIL 2006

Paul P. Acarnley received the B.Sc. and Ph.D. degrees in electrical engineering from Leeds University, Leeds, U.K., in 1974 and 1977, respectively, and the M.A. degree from Cambridge University, Cambridge, U.K., in 1978. After seven years in the Department of Engineering at Cambridge University, he joined the Power Electronics, Drives, and Machines Group at the University of Newcastle upon Tyne, U.K., in 1986. He authored Stepping Motors: A Guide to Theory and Practice, which was rst published by the Institution of Electrical Engineers, London, U. K., in 1982 with the fourth edition appearing in 2002. In 2003, he founded Research Engineering Education Services (RESEEDS) based in Stonehaven, U.K. He is also a Research Professor in the School of Engineering at Robert Gordon University (RGU), Aberdeen, U.K., and Emeritus Professor at the University of Newcastle upon Tyne, U.K. His principal research interest is in the control of electric drives, including work on state and parameter estimation applied to torque, current, temperature, speed, and position estimation in motors, and temperature estimation in power electronic devices. In addition, he has made contributions in the areas of stepping motors, permanent-magnet (PM) generators, ywheel energy storage, and brushless dc drives. Prof. Acarnley is a Fellow of the Institution of Electrical Engineers, U.K.

John F. Watson (SM94) was born in Kirkcaldy, U.K., in 1955. He received the degree in electronic and electrical engineering (with honors) from Heriot Watt University, Edinburgh, U.K., in 1973, and the Ph.D. degree in electrical engineering from the Council for National Academic Awards (CNAA) in 1981. He was a Development Engineer with GE (U.K.) for two years before joining the Robert Gordon University (RGU), Aberdeen, U.K., in 1982. He is currently a Professor and the Head of the School of Engineering at RGU. His main research interests are in the elds of condition monitoring and electrical drives. Prof. Watson is a Chartered Engineer in the U.K. and a Fellow of the Institution of Electrical Engineers, U.K.

Vous aimerez peut-être aussi