• No results found

Model-based commutation of a long-stroke magnetically levitated linear actuator

N/A
N/A
Protected

Academic year: 2021

Share "Model-based commutation of a long-stroke magnetically levitated linear actuator"

Copied!
10
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Model-based commutation of a long-stroke magnetically

levitated linear actuator

Citation for published version (APA):

Lierop, van, C. M. M., Jansen, J. W., Damen, A. A. H., Lomonova, E. A., Bosch, van den, P. P. J., & Vandenput, A. J. A. (2009). Model-based commutation of a long-stroke magnetically levitated linear actuator. IEEE

Transactions on Industry Applications, 45(6), 1982-1990. https://doi.org/10.1109/TIA.2009.2031783

DOI:

10.1109/TIA.2009.2031783

Document status and date: Published: 01/01/2009 Document Version:

Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication:

• A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website.

• The final author version and the galley proof are versions of the publication after peer review.

• The final published version features the final layout of the paper including the volume, issue and page numbers.

Link to publication

General rights

Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain

• You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement:

www.tue.nl/taverne

Take down policy

If you believe that this document breaches copyright please contact us at: openaccess@tue.nl

providing details and we will investigate your claim.

(2)

Model-Based Commutation of a Long-Stroke

Magnetically Levitated Linear Actuator

C. M. M. van Lierop, Member, IEEE, J. W. Jansen, Member, IEEE, A. A. H. Damen,

E. A. Lomonova, Senior Member, IEEE, P. P. J. van den Bosch, Member, IEEE, and

A. J. A. Vandenput, Senior Member, IEEE

Abstract—This paper concerns a commutation and

switch-ing algorithm for multi-degree-of-freedom (DOF) movswitch-ing-magnet actuators with permanent magnets and integrated active magnetic bearing. Because of the integration of long-stroke actuation and an active magnetic bearing, DQ-decomposition cannot be applied. Therefore, these actuators are a special class of synchronous ma-chines. A newly developed model-based commutation algorithm for linear and planar actuators enables long-stroke motion by switching between different sets of coils. Moreover, the ohmic losses in the coils are minimized. Three different versions of the algorithm are compared and successfully implemented on a three-DOF magnetically levitated linear actuator.

Index Terms—Commutation, linear synchronous motors, magnetic levitation, permanent-magnet motors.

I. INTRODUCTION

I

N RECENT YEARS, magnetically levitated planar actu-ators have been developed as an alternative to xy-drives constructed of stacked linear actuators. Although the translator of these actuators can move over relatively large distances in the xy-plane only, it has to be controlled in six degrees-of-freedom (DOFs) because of the active magnetic bearing. These actuators have either moving coils and stationary magnets [1] or moving magnets and stationary coils [2], [3]. The coils in the actuator are simultaneously used for propulsion in the xy-plane as well as for the four-DOF active magnetic bearing. To integrate the long-stroke propulsion and the active magnetic bearing of a moving-magnet actuator, new design [4] and control methods [5] have been developed.

Both planar actuator configurations enable long-stroke movement in the xy-plane. The main advantage of the moving-magnet configuration over the moving-coil configuration is that

Paper IPCSD-09-006, presented at the 2006 Industry Applications Society Annual Meeting, Tampa, FL, October 8–12, and approved for publication in the IEEE TRANSACTIONS ONINDUSTRYAPPLICATIONSby the Electric Ma-chines Committee of the IEEE Industry Applications Society. Manuscript sub-mitted for review February 1, 2007 and released for publication May 14, 2009. First published September 15, 2009; current version published November 18, 2009. This work (IOP-EMVT project) was supported by SenterNovem, an agency of the Dutch Ministry of Economical Affairs.

C. M. M. van Lierop, J. W. Jansen, A. A. H. Damen, E. A. Lomonova, and P. P. J. van den Bosch are with the Department of Electrical Engineering, Eindhoven University of Technology, 5600 MB Eindhoven, The Netherlands (e-mail: c.m.m.v.lierop@ieee.org; j.w.jansen@tue.nl; a.a.h.damen@tue.nl; e.lomonova@tue.nl; p.p.j.v.d.bosch@tue.nl).

A. J. A. Vandenput, deceased, was with the Department of Electrical Engineering, Eindhoven University of Technology, 5600 MB Eindhoven, The Netherlands (e-mail: a.j.a.vandenput@tue.nl).

Color versions of one or more of the figures in this paper are available online at http://ieeexplore.ieee.org.

Digital Object Identifier 10.1109/TIA.2009.2031783

the coils, and, therefore, the cooling and the power supply, are on the stationary part of the actuator. Consequently, there is no need for a cable to the translator. The stroke of the moving-magnet planar actuator is limited by the amount of stationary coils. Only the coils below the translator of the moving-magnet structure significantly contribute to its propulsion and its active magnetic bearing. Consequently, during the movement of the translator, the set of active coils needs to be switched. The decoupling of the DOFs that can be achieved by optimization of the moving-magnet actuator is fundamentally limited when using standard DQ-decomposition, since the design symmetries needed for decoupling do not exist [5] (as opposed to the moving-coil topology [1]). Due to the fundamentally larger coupling between the DOFs in combination with the need to switch between active coils, Park-transformation (DQ0-decomposition) and vector control cannot be applied. There-fore, a nonlinear model-based commutation algorithm has been derived [5] which controls each coil individually. The algorithm enables long-stroke motion, without compromising the decou-pling of the DOFs.

This paper provides an implementation of the new com-mutation algorithm on a long-stroke three-DOF magnetically levitated linear actuator [4]. A simplified real-time model [6] of the three-DOF actuator has been used for model-based control. A comparison of three implementations of the model-based commutation has been made using different sizes of active coil sets. The three-DOF linear actuator incorporates the same integration of propulsion and active magnetic bearing as a six-DOF long-stroke moving-magnet planar actuator, which is the final goal of this paper.

II. TOPOLOGY

The bottom and side views of the three-DOF actuator are shown in Fig. 1. The three DOFs, propulsion in the x-direction, levitation in the z-direction, and rotation of the translator about the y-axis, are indicated in the side view. The translator is a Halbach magnet array. The magnets in the array are rotated 45 with respect to the coils. The stator consists of ten coils. Five adjacent coils are simultaneously energized to control the three DOFs. The total stroke in the x-direction is approximately 191 mm. The distance between the coils is 3/2τn. The

defini-tion of the pole pitch τn is shown in Fig. 1. The vector cp =

[cpxcpz]indicates the position of the translator with respect to

the origin of the global coil-coordinate system. The actuator can be classified as an ac-synchronous machine. However, due to

(3)

Fig. 1. (a) Bottom view of the actuator. (b) Side view of the actuator.

Fig. 2. (a) Bottom and (b) side views of a solid coil located atcx = 0,cy = 0, cz = 0, an equivalent filament coil, and the magnet array with the coordinate

systems.

the nonsinusoidal commutation, it is not a polyphase actuator. Therefore, each coil is connected to a single-phase amplifier.

III. MODEL

A. Single-Coil Real-Time Model

To derive a model-based commutation algorithm, an analyti-cal real-time model of the three-DOF actuator is required. The model which is used to derive the forces and torque of one coil is presented in [6]. The force and torque on the translator can be calculated with the Lorentz force method. However, the force on the magnets is opposite to the force on the coils. To reduce the model complexity, the volume integral of the Lorentz force equation can be reduced to a closed line integral by modeling the coil as four filaments as shown in Fig. 2. The force

cF

transand the torquecTtranson the translator, expressed in the

global coil-coordinate system (defined in Fig. 1 and indicated by superscriptc), then become

cF trans=  C ci f×cBdl (1) cT trans=  C cr×ci f ×cB  dl (2) where ci

f is the filament current vector in ampere-turns, cB

is a magnetic-flux-density distribution in coil coordinates, and

cr is the distance to the mass center point. The

magnetic-flux-density distributionmB is simplified to a 2-D sine wave defined in the local magnet coordinatesmx andmy, as shown in Fig. 2,

and is then transformed to the global coil-coordinate system

cB (as defined in Fig. 1). The error which is introduced by

neglecting the higher space harmonics of the magnetic-flux-density distributioncB is small, since the design is optimized to minimize these harmonics [4]. Assuming the translator angle θ = 0 rad, the force and torque on the mass center point of the translator then become [6]

cF trans = [cF x cFy cFz] = ⎡ ⎢ ⎣ N 4Bzτne−λ cp zsin  π(cp x−cx) τn  i 0 N 4√2Bxyτne−λ cp zcos  π(cp x−cx) τn  i ⎤ ⎥ ⎦ (3) cT trans = [cT x cTy cTz] = ⎡ ⎢ ⎣ N 2√2Bxyτn2e−λpzsin  π(cp y−cy) τ  i−(cp y−cy)cFz (cp x−cx)cFz−cpczFx (cp y−cy)cFx ⎤ ⎥ ⎦ (4) where (cx,cy, 0) is the position of the center of the coil in

global coil coordinates as defined in Fig. 2,cp

xandcpzindicate

the translator position in global coil coordinates as defined in Fig. 1, i is the coil current in amperes, τn is the new effective

pole pitch after rotation of the magnet array, λ = π/τn is a

constant representing the decline of the magnetic flux density, N is the number of turns, and Bxy and Bz are the effective

amplitudes of the sinusoidal flux density atmz = 0. It is also possible to include the angle θ in the model [6]. For simplicity, the force and torque equations are rewritten into

 w = [cF

x cFz cTy]= γ (cpx,cpz,cx) i (5)

where the wrench vector w is a vector containing the force and torque components which are nonzero after assuming cpy= cy = 0, andcx is the location of the center of the coil along

the x-axis of the global coil-coordinate system. B. Full Three-DOF Actuator Model

So far, the influence of only one coil has been considered. The total wrench exerted on the translator is calculated using superposition of each individual coil, assuming rigid-body dy-namics. After superposition, the wrench vector w can be ob-tained by multiplying a [3× 10] position-dependent matrixcΓc

by a current vector i = [ i1 i2 · · · i10] which contains

the individual currents of the ten coils, as shown by (6) at the bottom of the next page. This model does not contain the end effects of the magnet array. Therefore, it is only valid for the coils which are not near or outside the periphery of the magnet

(4)

array. This does not present a big disadvantage, since only the coils near the center of the translator will be excited.

C. Simplified Three-DOF Actuator Model

The three-DOF actuator is controlled using only a subset of the ten coils. The smallest simplified model contains at least five simultaneously energized adjacent coils. Due to the repeating pattern of the stationary coils in the x-direction, only five coils need to be modeled simultaneously. Therefore, a simplified model is proposed, as shown by (7) at the bot-tom of the page, where cΓ

r is a [3× 5] position-dependent

matrix, rp

x∈ [0, 3/2τn) is a repeating coordinate along the

x-axis (the period of repetition is equal to the distance between the coils of 3/2τn), rpz=cpx, and iα is a vector

contain-ing the five active currents. An absolute-position counter α∈ {−3, −2, −1, 0, 1, 2} is introduced, which function is twofold. First, the counter can be used to select the proper origin of the repeating coordinate system (indicated by superscript r).

The origin is shifted by 3/2ατn along the x-axis with

re-spect to the global coil-coordinate system, e.g., when α = 0, the origin of the relative position coordinates is located between coil 5 and coil 6 which is also the origin of the global coil-coordinate system (which is shown in Fig. 1), or when α = 1, the origin of the relative position coordinates is shifted by 3/2τn along the x-axis (located between coil

6 and coil 7, which is shown in Fig. 3), etc. Second, it can be used to select the proper active current vector iα, e.g.,

when α = 0, iα= [ i4 i5 i6 i7 i8], or when α = 1,

iα= [ i5 i6 i7 i8 i9], etc. It is also possible to use

more coils simultaneously. However, the unmodeled edge-effect of the magnet array will reduce the model accuracy of

Fig. 3. Side view of the actuator including the coil-current weighing matrix

cΔ

r(rpx) when using five active currents.

the coils located near the edges of the magnet array. In order to identify the influence of the edge effect on the model-based commutation algorithm, two more models will be considered: a model using six active coils, as shown by (8) at the bottom of the page, in which casecΓ

rbecomes a [3× 6] matrix with α ∈

{−2, −1, 0, 1, 2} andrp

x∈ [−3/4τn, 3/4τn), as shown at the

bottom of the page, and a model using seven active coils, shown by (9) at the bottom of the page, in which casecΓ

rbecomes a

[3× 7] matrix with α ∈ {−2, −1, 0, 1} andrp

x∈ [0, 3/2τn),

as shown at the bottom of the page. The inaccuracy introduced by the edge effect can be reduced by applying correction terms to the model of the coils located near the edge of the magnet array. It is also possible to use data-based methods to derive the correct values of the wrench as a function of the currents through the coils in proximity of the edges of the magnet array. However, this is beyond the scope of this paper.

IV. COMMUTATION ANDSWITCHING

A. Short-Stroke Direct Wrench-Current Decoupling

Since the force distribution over the translator has to be taken into account, each active coil has to be controlled individually. Therefore, linearization by feedback is used to linearize and

 w =cΓc(cpx,cpz)i = [ γ cp x,cpz,−92 2n  γ cp x,cpz,−72 2n · · · γ cp x,cpz,922n ]i (6)  w =cΓr(rpx,rpz)iα = [ γ rpx,rpz,−32 2n  γ rpx,rpz,−12 2n · · · γ rp x,rpz,522n ]iα (7)  w =cΓr(rpx,rpz)iα = [ γ rp x,rpz,−52 2n  γ rp x,rpz,−32 2n · · · γ rp x,rpz,522n ]iα (8)  w =cΓr(rpx,rpz)iα = [ γ rp x,rpz,−52 2n  γ rp x,rpz,−32 2n · · · γ rp x,rpz,722n ]iα (9)

(5)

decouple the DOFs, instead of using DQ0-transformation [5]. The advantage of this approach, when compared with the DQ0-approach, is that it results in a decoupling of both the forces and the torque. The three-DOF actuator is clearly overactuated (there are more active coils than DOFs). Consequently, there is a set of transformations that lead to a proper linearization and decoupling of the DOFs. Therefore, the ohmic losses in the coil can be minimized leading to the following optimiza-tion problem:

min

cΓr(rpx,rpz)iα= wiα = cΓ

r(rpx,rpz) w (10)

where the two-norm of the active current vector iα, i.e., the

ohmic losses in the coils, is minimized while maintaining the relation between the current and the wrench vector given by (7). Equation (10) can be solved by using the generalized inverse of a minimum-norm solution of a consistent equation [7], resulting in cΓ r(rpx,rpz) =r(rpx,rpz) c Γr(rpx,rpz)r(rpx,rpz) −1 . (11) B. Long-Stroke Direct Wrench-Current Decoupling

In order to achieve long-stroke movement, it is necessary to switch between active coils. Therefore, (11) only presents a solution to the minimization of the ohmic losses in the coils (10) for short-stroke motion, since it does not incorporate any switching constraints.

Due to the finite voltage specification of the amplifiers which drive the individual coils, the time derivative of the currents through the coils is limited. Therefore, an additional constraint has to be added to ensure smooth switching of the coils. A possible strategy can be to switch a coil off when the cur-rent through that coil becomes zero. However, this does not guarantee that the coil can be switched off before it leaves the surface of the translator, at which point, the model is not valid anymore (as explained in Section III-C). When there is a priori information about the trajectory, it is possible to check whether (11) can be used for that trajectory, i.e., the trajectory in combination with (11) should always result in a zero crossing of the currents through the coils that need to be switched off. However, this leads to a difficult and very time consum-ing optimization (e.g., mixed-integer quadratic programmconsum-ing) which has to be calculated offline for each trajectory. A more practical solution, which does not need a priori information about the trajectory, is to actively force the currents through the coils near the edge of the translator to zero by using smooth position-dependent weighing functions which penalize each current.

When minimizing the ohmic losses, including the switching constraints, the following optimization needs to be solved:

min

cΓ

r(rpx,rpz)iα= w

iαΔ−1(rpx)= r(rpx,rpz) w (12) where the weighted two-norm of the vector iα is minimized,

using a current-vector weighing matrixcΔ

r(rpx) while

main-taining the relation between the current-and the wrench-vector

given by (7). When using five active currents, the weighing matrixcΔ

r(rpx) consists of the following structure:

cΔ r(rpx) = ⎡ ⎢ ⎢ ⎢ ⎣ cf r,wl(rpx) 0 0 0 0 0 1 0 0 0 0 0 1 0 0 0 0 0 1 0 0 0 0 0 cf r,wr(rpx) ⎤ ⎥ ⎥ ⎥ ⎦ (13) with cf r,wl(rpx) = 1 2+ 1 2cos  3τn rp x  (14) cf r,wr(rpx) = 1 2 1 2cos  3τn rp x  (15) where the elements on the diagonal of the matrix represent a penalty to each individual active coil current. If the elements are equal to one, there is no penalty, and if they are equal to zero, the corresponding active coil current is forced to zero. Since the active set of coil currents must be smoothly switched from one set to another to achieve long-stroke motion, the outermost coils of the active coil-current set need to be penalized. As can be seen from (13), the first and last elements contain functions which are dependent on the repeating x-position rpx [(14),

(15)], which is illustrated in Fig. 3. Naturally, the same can be done for six or seven active coils.

Again, (12) can be solved by using the generalized inverse of a minimum-norm solution of a consistent equation [7], resulting in cΓ r(rpx,rpz) = cΔ r(rpx)cΓr(rpx,rpz) c Γr(rpx,rpz)cΔr(rpx)r(rpx,rpz) −1 . (16)

C. Reduction of Calculation Time

Equation (16) has been implemented using two different strategies. The most straightforward implementation strategy is to derive the analytical solution of (16) and use the solution to calculate the position-dependent currents. A second imple-mentation method is to perform a real-time numerical matrix inverse operation. The amount of calculation time needed to solve all trigonometric operations of the analytical solution is much longer than the amount of calculation time needed for the numerical real-time matrix inverse. Moreover, the implementa-tion method using a real-time matrix inverse enables easy offset correction of the coil locations.

V. SIMULATION ANDEXPERIMENTALVERIFICATION

A three-DOF moving-magnet linear preprototype which in-corporates the same integration of propulsion and active mag-netic bearing as a six-DOF moving-magnet planar actuator has been designed, optimized, and manufactured [4]. The sizes of the actuator are summarized in Table I. Figs. 4 and 5 show a photograph of the stator and of the translator, respectively.

To test this three-DOF actuator, the translator is attached to a mechanical H-drive with three coupled linear motors. The

(6)

TABLE I

DIMENSIONS OF THEACTUATOR

Fig. 4. Stator.

Fig. 5. Translator.

Fig. 6. Setup to test the commutation.

H-drive (or gantry) allows positioning of the translator of the three-DOF actuator in three DOFs with respect to the stator coils. For this experiment, the three-DOF actuator is rotated 90 about its x-axis, i.e., put on its side. Due to this rotation, the DOFs of both systems match. A schematic top view and a photograph of the three-DOF actuator and the H-drive are shown in Figs. 6 and 7(a), respectively. A six-DOF load cell Fig. 7(b) (JR3 45E15A4-I63-S 100N10) is mounted between

Fig. 7. (a) Photograph of the setup. (b) Side view with load cell.

Fig. 8. Simulated current waveforms using five active coils,cF

x= 10 N, cF

z= 10 N,cTy= 0 Nm, a mechanical clearance of 2 mm, and θ = 0 rad.

The areas in which the different coil sets are active are indicated with the absolute-position counter α.

the translator of the H-drive and the translator of the three-DOF actuator to measure the forces and torques on the translator. The commutation and switching algorithm has been tested in open loop.

While the H-drive moves the translator in the x-direction along the stator coils at 0.02 m/s, with a mechanical clearance of 2 mm and θ = 0 rad, a constant wrench of cF

x= 10 N, cF

z= 10 N, and cTy = 0 Nm is produced on the translator

by excitation of the stator coils. Every stator coil has been connected to a single-phase current amplifier. The current waveforms of the commutation using five, six, and seven active coil currents are shown in Figs. 8–10, respectively. The current waveforms are nonsinusoidal, but they repeat in time.

Because the commutation is measured in open loop, the result is very sensitive to power-amplifier offsets, variation in the magnet properties, and tolerances in the exact coil locations. Therefore, all power-amplifier offsets and coil locations have been individually adjusted in the algorithm. This has been done by measuring the force produced by a single coil at a constant current. According to (3), the force along the x-axis

cF

x depends sinusoidally on the difference between the coil

offsetcx and the translator positioncp

x. Therefore, when the

coil offset is correct, the measured force is zero forcp x=cx.

(7)

Fig. 9. Simulated current waveforms using six active coils,cF

x= 10 N, cF

z= 10 N,cTy= 0 Nm, a mechanical clearance of 2 mm, and θ = 0 rad.

The areas in which the different coil sets are active are indicated with the absolute-position counter α.

Fig. 10. Simulated current waveforms using seven active coils,cFx= 10 N, cF

z= 10 N,cTy= 0 Nm, a mechanical clearance of 2 mm and θ = 0 rad.

The areas in which the different coil sets are active are indicated with the absolute-position counter α.

The individual coil-offset adjustments result in small changes of the current waveforms, which is shown in Figs. 11 and 12.

Figs. 13–18 are three sets of two figures showing the predicted and measured force and torque waveforms of the commutation using five, six and seven active coil currents, respectively. The measured data of the forces cF

x, cFz and

torquecT

yas a function of time have been filtered offline using

an eighth-order anticausal Butterworth filter with a bandwidth of 25 Hz. Apart from an increased force and torque ripple, the measurements are in good agreement with the predicted values. The measured force ripple is close to the measurement accuracy of the load cell. Moreover, the predicted force and torque ripple is calculated using ideal coil and magnet dimen-sions. The measured ripple using the commutation algorithm is minimized by adjusting the individual coil and amplifier offsets. However, the amount of harmonic disturbance of each individual coil, resulting from the tolerances in the magnet and coil dimensions, leads to a higher commutated force and torque ripple. Figs. 13–18 clearly illustrate that the desired

Fig. 11. Current waveforms using five active coils,cF

x= 10 N,cFz=

10 N,cT

y= 0 Nm, a mechanical clearance of 2 mm, and θ = 0 rad after

coil offsetcx correction. The areas in which the different coil sets are active are

indicated with the absolute-position counter α.

Fig. 12. Error between the uncorrected and the corrected current waveforms shown in Figs. 8 and 11, respectively.

Fig. 13. Predicted and measured open-loop forces using five active coils during commutationcF

x= 10 N,cFz= 10 N,cTy= 0 Nm, a mechanical

(8)

Fig. 14. Predicted and measured open-loop torque using five active coils during commutationcF

x= 10 N,cFz= 10 N,cTy= 0 Nm, a mechanical

clearance of 2 mm, and θ = 0 rad.

Fig. 15. Predicted and measured open-loop forces using six active coils during commutationcFx= 10 N,cFz= 10 N,cTy= 0 Nm, a mechanical clearance

of 2 mm, and θ = 0 rad.

Fig. 16. Predicted and measured open-loop torque using six active coils during commutationcF

x= 10 N,cFz= 10 N,cTy= 0 Nm, a mechanical

clearance of 2 mm, and θ = 0 rad.

Fig. 17. Predicted and measured open-loop forces using seven active coils during commutationcF

x= 10 N,cFz= 10 N,cTy= 0 Nm, a mechanical

clearance of 2 mm, and θ = 0 rad.

Fig. 18. Predicted and measured open-loop torque using seven active coils during commutationcF

x= 10 N,cFz= 10 N,cTy= 0 Nm, a mechanical

clearance of 2 mm, and θ = 0 rad.

values of the reference forces and torques are obtained. The repeating patterns also coincide with the amount of switching moments (six periods when using five active coils, five periods when using six active coils, and four periods when using seven active coils). Increasing the amount of active currents beyond the minimum of five (which is necessary for controllability [5]) reduces not only the power dissipation but also the stroke. The higher measured force and torque ripple still contains the same periodic behavior as the switching moments because of the position-dependent sensitivity to model disturbances of the commutation algorithm. This position-dependent sensitivity is caused by the switching between active coils as well as the changing force distribution over the translator as a function of position [8]. When using seven active coils, the model of the outer coils which need to be switched on or off is not accurate anymore. However, since these coils are heavily penalized, the model errors of these coils do not significantly contribute to the total decoupling of the commutation algorithm. Errors in the total decoupling occur when the amount of simultaneously

(9)

active coils is increased beyond seven. These errors are a result of the influence of the unmodeled-edge effects of the magnet array on the active coils which are not yet penalized by the window functions. The measurement results are in good agree-ment with the simulation results and, therefore, elucidate the good performance of the model-based commutation strategy.

The success of the experiments on the open-loop behavior of the direct wrench-current decoupling resulted in the application of the new commutation algorithm onto a six-DOF planar actuator with moving magnets [8], [9]. The tracking error of this planar actuator is 30 μm and 100 μrad at full speed (1.4 m/s) and full acceleration (14 m/s2) [10].

VI. CONCLUSION

A new commutation algorithm, which enables combined long-stroke propulsion and active magnetic-bearing control of an ironless multi-DOF moving-magnet actuator, including smooth switching of the currents through the active coil sets, has been successfully implemented. The algorithm is model based and makes use of the repeating coil-structure of the moving-magnet actuator. Measurements show that the overall error on the model-based commutation caused by neglecting the edge effect is small, as long as the active coils located near the edges of the magnet array are properly penalized in the commutation algorithm. The measurement results are in good agreement with the simulation results and, therefore, elu-cidate the good performance of the model-based commutation strategy.

ACKNOWLEDGMENT

The authors would like to thank M. J. P. C. Uyt De Willigen for his assistance with the experimental setup and also the companies, ASML, Prodrive, Tecnotion, and Assembléon, for their valuable contributions and discussions.

REFERENCES

[1] J. C. Compter, “Electro-dynamic planar motor,” Precis. Eng., vol. 28, no. 2, pp. 171–180, Apr. 2004.

[2] W. J. Kim, D. L. Trumper, and J. H. Lang, “Modeling and vector control of a planar magnetic levitator,” IEEE Trans. Ind. Appl., vol. 34, no. 6, pp. 1254–1262, Nov./Dec. 1998.

[3] A. J. Hazelton, M. B. Binnard, and J. M. Gery, “Electric motors and positioning devices having moving magnet arrays and six degrees of freedom,” U.S. Patent 6 208 045, Mar. 27, 2001.

[4] J. W. Jansen, J. M. van Essen, C. M. M. van Lierop, E. A. Lomonova, and A. J. A. Vandenput, “Design and test of an ironless, three degree-of-freedom, magnetically levitated linear actuator with moving magnets,” in

Proc. IEEE IEMDC, Jun. 2005, pp. 93–96.

[5] C. M. M. van Lierop, J. W. Jansen, A. A. H. Damen, and P. P. J. van den Bosch, “Control of multi-degree-of-freedom planar actuators,” in Proc. IEEE Int. CCA, Munich, Germany, Oct. 2006, pp. 2516–2521.

[6] J. W. Jansen, C. M. M. van Lierop, E. A. Lomonova, and A. J. A. Vandenput, “Modeling of a magnetically levitated planar actuator with moving magnets,” IEEE Trans. Magn., vol. 43, no. 1, pp. 15–25, Jan. 2007.

[7] C. R. Rao and S. K. Mitra, Generalized Inverse of Matrices and Its

Applications. New York: Wiley, 1971.

[8] C. M. M. van Lierop, “Magnetically levitated planar actuator with moving magnets: Dynamics, commutation and control design,” Ph.D. dissertation, Eindhoven Univ. Technol., Eindhoven, The Netherlands, Jan. 2008.

[9] J. W. Jansen, “Magnetically levitated planar actuator with moving magnets: Electromechanical analysis and design,” Ph.D. dissertation, Eindhoven Univ. Technol., Eindhoven, The Netherlands, Nov. 2007. [10] J. W. Jansen, C. M. M. van Lierop, E. A. Lomonova, and

A. J. A. Vandenput, “Magnetically levitated planar actuator with mov-ing magnets,” IEEE Trans. Ind. Appl., vol. 44, no. 4, pp. 1108–1115, Jul./Aug. 2008.

C. M. M. van Lierop (S’05–M’07) was born in

Eindhoven, The Netherlands. He received the M.Sc. degree (cum laude) in electrical engineering and the Ph.D. degree in magnetically levitated planar actuator technology from Eindhoven University of Technology, Eindhoven, in 2003 and 2008, respectively.

He has been with the Control Systems Group, Eindhoven University of Technology, where he was initially a Postdoctoral Researcher and is currently an Assistant Professor. His main research interests deal with motion control, magnetic bearings, electron microscopy, and MIMO systems.

J. W. Jansen (S’03–M’07) received the M.Sc.

degree (cum laude) in electrical engineering and the Ph.D. degree in magnetically levitated planar actuator technology from Eindhoven University of Technology, Eindhoven, The Netherlands, in 2003 and 2007, respectively.

He has been with the Electromechanics and Power Electronics Group, Eindhoven University of Tech-nology, where he was initially a Postdoctoral Re-searcher and is currently an Assistant Professor. His research interests include magnetic levitation and the analysis and design of linear and planar actuators.

A. A. H. Damen was born in ’s-Hertogenbosch,

The Netherlands, in 1946. He received the M.Sc. degree in electrical engineering from the Measure-ment and Control Group, Eindhoven University of Technology, Eindhoven, The Netherlands, in 1970, and the Ph.D. degree based on the thesis “On the Observability of Electrical Cardiac Sources,” in 1980 at the same university.

Since 1979, he has been an Associate Professor in the Measurement and Control Group, Eindhoven University of Technology. After his analysis of the inverse problem in cardiography, he contributed to the field of system identification, and, at the moment, his main interest lies in robust control.

E. A. Lomonova (M’04–SM’07) was born in

Moscow, Russia. She received the M.Sc. (cum laude) and Ph.D. (cum laude) degrees in electromechan-ical engineering from the Moscow State Aviation Institute (TU), Moscow, Russia, in 1982 and 1993, respectively.

She has been with Eindhoven University of Tech-nology, Eindhoven, The Netherlands, where she was an Assistant Professor in 2001 and has been an Associate Professor since 2008. She has worked on electromechanical actuator design, optimization, and development of advanced mechatronic systems.

(10)

P. P. J. van den Bosch (M’84) received the M.Sc.

degree (cum laude) in electrical engineering and the Ph.D. degree in optimization of electric energy systems from Delft University of Technology, Delft, The Netherlands.

After his studies, he joined the Control Sys-tems Group, Delft University of Technology, and was appointed Full Professor in control engineering, in 1988. He has been with the Control Systems Group, Electrical Engineering, Eindhoven Univer-sity of Technology, Eindhoven, The Netherlands, where, in 1993, he was appointed Full Professor, and in 2004, he was ap-pointed a part-time Professor in the Department of Biomedical Engineering. His research interests include modeling, optimization, and control of dynam-ical systems. In his career, he has created many common research projects with industry: automotive applications, large-scale electric systems, advanced electromechanical actuators, and, recently, embedded systems and biomedical modeling, among others.

Dr. van den Bosch has been a recipient of several prizes for his educational activities and several patents. He has served on the Boards of journals (Journal A) and conference committees.

A. J. A. Vandenput (M’77–SM’87) received the

Ph.D. degree from the Catholic University of Leuven, Leuven, Belgium, in 1977.

From 1973 to 1991, he was with the Laboratory for Electrical Machines and Drives, Catholic Univer-sity of Leuven, first as a Researcher of the Belgian National Fund for Scientific Research and, in 1987, as a Professor. He obtained an Alexander von Humboldt-Stipendium for performing research at the RWTH Aachen, Aachen, Germany, from 1980 to 1981; was a Visiting Professor at the University of Colorado, Boulder, from 1983 to 1984; and he was a Research Fellow with the Japan Society for the Promotion of Science (JSPS), University of Tokyo, Tokyo, Japan, in 1986. From 1991 to 2008, he was a full-time Professor and Chairman of the Electromechanics and Power Electronics Section, Eindhoven Univer-sity of Technology, Eindhoven, The Netherlands. His teaching and research interests were high-performance electric drives, including electrical machines, power electronics, and control techniques. He passed away on January 11, 2008.

Referenties

GERELATEERDE DOCUMENTEN

Moreover, the commutation algorithm also allows for smooth switching of active coil sets without influencing the decoupling, because only the coils that are underneath or near the

principle can be applied to the force and torque production, the sizes of the magnets and coils in a planar actuator are determined by only optimizing the sizes of the magnet array

Whereas other moving- magnet planar actuators 2 , 4 require a redesign of the actuator, the stroke of the investigated planar actuator can be in- creased by simply adding extra

If coils effected by the end effect of the magnet array (edge coils) are added to the active coil set, an error will occur in the commutation because the harmonic model is

The effect of the gas-liquid mass transfer on the overall reaction rate was separately determined in a three-phase bubble column using a rapid hydrogenation

In addition to the biological relationship, the ion images also show a similar spatial distribution to the segmentation obtained from the hyperspectral visualizations, as shown

It is often of great interest to extract the common information present in a given set of signals or multichannel recording. This is possible by going to the subspace representation

In conclusion, our results indicate that the experimental loop design, together with a statistical analysis by a linear model, provides an efficient procedure for the detection