Sunteți pe pagina 1din 582

GNA PHYSICS PUBLICATIONS OXFORD SCIENCE PUBLICATI

DESIGN OF BRUSHLESS
PERMANENT-MAGNET
MOTORS
j. R. HENDERSHOT J r
and TJE MILLER
MONOGRAPHS IN ELECTRICAL AND ELECTRONIC ENGINEERING

ftruh!e*5 permanem-mtfgnet moU*rs provide mmplc. law inpintemmcc, easily


controlled motive power New magnetic materials and digital power control
techniques continue to widen the scope of their applications: Increasing automa
tion is increasing demand.
This book, written by two world-leading experts. o.imbiiie:, practical and aca
demic experience to produce the most comprcheniuvc and useful design manual
for hrush less permanent-niaghei motors ever It is dcnigncd to se n t the modern
computer based generation of mo;or engineers, '['he book. uoe hand-in-hand
with modern software-based tedwiquer- for design und analysi*. Ij coven nil the
topic* of interest to the motor design engineer and the systems engineer, incjtid-
mg electrical and magnetic design, materials. control, and man> other topics.
J. R. Hendershot ,|r is General Manager of the Magna Physic* Hivision of
Trideha Industries. Hillsboro, Ohlu TJli Miller in I ,ucan Professor of Power
Engineering, University of Glasgow und Director of the SPEi) Consortium
MONOGRAPHS IN KUiCTRICAI AND ELECTRONIC ENGINEERING
Series editors. P ll-inimond. TJE Miller, S Yamamura
Brushless permancnt-m agnet and reluctance m otor drives (I9K9)
TJE Miller
Switched reluctance m otors anti their control (1993)
T J E M ille r
Numerical modelling of eddy currents (1993)
Andrzej Kruwczyk and John A Tegopoulos
Rectifiers, cycloconvertcrs. and a.c. controllers (1994)
Thornas H. Barton
Stepping m otors and their m icroprocessor controls: Second edition (1994)
T. Kcmo und A Sugawaru
Inverse problem s and optimal design in electricity and m agnetism (1994)
P. Neittaanmaki, M Rudnmki. and A. Savim

Magna Physics Publications i* u division of Tridclta Industries and specializes in


books on electric motors- actuators. and controls Which focus on and emphasize
the acfual design techniques more limn the tlienreucal principles of operation.
Monographs in Electrical and Electronic Engineering

10. The theory of linear Liduction machinery (1980) Michel Poloujadoff


12. Energy methods in electromagnetism (1981) P. Hammond
15. Superconducting rotating electrical machines (1983) J. R. Bumby
16. Stepping motors and their microprocessor controls (1984) T. Kenjo
17. Machinery noise measurement (1985) S. J. Yang and A. J. Ellison
18. Permanent-magnet and brushless d.c. motors (1985) T. Kenjo and S. Nagamori
19. Metal-semiconductor contracts: Second edition (1988) E. H. Rhoderick and
R. H. Williams
20. Introduction to power electronics (1988) Eiichi Ohno
21. Brushless permanent-magnet and reluctance motor drives (1989) TJE Miller
22. Vector control ofa.c. machines (1990) Peter Vas
23. Brushless servomotors: fundamentals and applications (1990) Y. Dote and
S. Kinoshita
24. Semiconductor devices, circuits, and systems (1991) Albrecht Moschwitzer
25. Electrical machines and drives: a space-vector theory approach (1992) Peter Vas
26. Spiral vector theory of a.c. circuits and machines (1992) Sakae Yamamura
27. Parameter estimation, condition monitoring, and diagnosis of electrical machines
(1993) Peter Vas
28. An introduction to ultrasonic motors (1993) S. Sashida and T. Kenjo
29. Ultrasonic motors: theory and applications (1993) S. Ueha and Y. Tomikawa
30. Linear induction drives (1993) J. F. Gieras
31. Switched reluctance motors and their control (1993) TJE Miller
32. Numerical modelling of eddy currents (1993) Andrzej Krawczyk and John
A. Tegopoulos
33. Rectifiers, cycloconverters, and a.c. controllers (1994) Thomas H. Barton
34. Stepping motors arui their microprocessor controls: Second edition (1994)
T. Kenjo and A. Sugawara
35. Inverse problems and optimal design in electricity and magnetism (1994)
P.Neittaanmaki, M. Rudnicki, and A. Savini
36. Applications of electric drive systems (1994) Richard M. Crowder
37. Design of Brushless Permanent-Magnet Motors (1994) J. R. Henderehot Jr,
and TJE Miller
Design of Brushless
Permanent-Magnet Motors

J.R. Hendershot Jr. TJE Miller


General Manager Lucas Professor in Power Electronics
Magna Physics Tridalta Director, SPEED Consortium
Hillsboro, OHIO 45133 University of Glasgow, UK

MAGNA PHYSICS PUBLISHING


AND
CLARENDON PRESS OXFORD
1994
Magna Physics Publishing, Hillsboro, OH 45133
Magna Physics Div. Tridclta Industries Inc., Mentor, OH 44060
and
Oxford University Press, Walton Street, Oxford 0X2 6DP
Oxford New York Toronto
Delhi Bombay Calcutta Madras Karachi
Kuala Lumpur Singapore Hong Kong Tokyo
Nairobi Dir es Salaam Cape Town
Melbourne Auckland Madrid
and associated companies in
' Berlin Ibadan
Oxford is a trade mark of Oxford University Press
Published in the United States
by Magna Physics Div. Tridelta Industries Inc., Hillsboro, Ohio
and Oxford University Press Inc., New York
J. R. Hendershot Jr. and T. J. E Miller, 1994
All rights reserved. No part of this publication may be
reproduced, stored in a retrieval system, or transmitted, in
any form of by any means, without the prior permission in writing
of Magna Physics or Oxford University Press. Within the USA & UK,
exceptions are allowed in respect of any fair dealing for the purpose of research or private
study, or criticism or review, as permitted under the Copyright, Designs and Patents Act,
1938, or in the ease of reprographic reproduction in accordance with the terras of licenses
issued by the Copyright Licensing Agency. Enquiries concerning reproduction
outside those terms and in other countries should be sent
to Magna Physics or Oxford University Press,
at the addresses shown above.
Library of Congress Cataloging in Publication Data is Available
A catalogue record for this book is available from the British Library
Magna Physics ISBN 1-881855-03-1
OUP ISBN 0-19-859389-9
Typeset by the Authors
Printed in (he United States of America
by Book Craftcrs, Chelsea, Michigan 48118-0370
Preface
Our purpose is to provide a straightforward approach to the practical
design of permanent-magnet brushless motors, supported by a detailed
account of many specialist topics. The practices, methods, theories and
calculations outlined herein are the result of many years experience in
design of all types of brushless machines, coupled with close cooperation
between the industrial manufacturing company and the academic
laboratory. We have tried to achieve a unity between theory and practice
that is the essence of good engineering.
The precursor to this book (written and published by JRH) has been
included almost in its entirety, with only minor revisions. Chapters on
specialist topics have been added by staff of the SPEED Laboratory at
Glasgow University,1 more than trebling the volume and covering the
magnetic and electrical design, the EMF waveform, core losses, electronic
commutation, computer-aided design, thermal analysis, control, materials,
field-weakening, and axially-laminated motors.
The first three descriptive chapters cover the design features and
application characteristics in broad terms, dealing with many of the
principal choices which must be made before a detailed design is
undertaken. This material should be easily understood by application
engineers, students, and others with no special experience in motor
design. The various brushless motor configurations are summarized, with
the pros and cons of each stated so that a proper selection may be
made at the outset of a design for a particular application.
Chapters 4-7 cover magnetic and electrical design in detail, presenting
many design equations simple enough for a calculator or spreadsheet
program. Chapter 13 gives detailed examples. Chapters 8-12 and 14-18
contain detailed analysis of special topics in design and performance.
Considerable effort has been made to distinguish between squarewave and
sinewave motors, because it is believed that the fundamental differences
between them (and their controllers) are not widely understood.

1
Dr. W.L. Soong is now with GE Corporate Research and Development, Schencctady, NY,
USA, and Dr. R. Rabinovici is at Ben-Gurion University of the Negev. Beer-Sheva, Israel.
P refa ce
There are so many independent variables involved in the design process,
that the reiteration method is required. By that we mean making certain
assumptions, assigning trial values to independent variables, and
calculating the dependent variables. The independent or "input"
variables are usually dimensions, winding turns, and magnet material
properties. The dependent or "output" variables are usually perform ance
figures such as torque, current, efficiency, temperature rise, etc. The
ensemble of output data is the "performance". If the performance is not
satisfactory, new estimates are made for the input variables, taking
account of trends observed in successive calculations. The process
continues until the desired performance is achieved.
The individual steps in the design process are essentially simple, but a
comprehensive design exercise may include hundreds of parameters and
thousands of iterations, and is clearly capable of being speeded up and
generally improved with the help o f a computer program such as the
SPEED Laboratorys PC-BDC program. This book follows the conventions
and equations used in PO-BDC, and in fact it can be used as a design
guide in conjunction with PC-BDC or, for that matter, with any equivalent
GAD software or design procedures.
We would like to thank everyone who has helped us, both direcdy and
indirecdy, including all those who provided material in the form of
photographs and diagrams, our customers, and especially our colleagues
and families. We would particularly like to acknowledge R.C. Perrine,
Gene Aha, and Malcolm McGilp on the engineering side, and Susie
Murdoch and Bridget Sweeney for their help with the manuscript.

J.R. Hendershot Jr. TJE Miller


Hillsboro, Ohio Glasgow, UK
June 1994
Contents
1. GENERAL INTRODUCTION
J.R Hendershot Jr.
1.1 Definitions and types of brushless m o to r................................... 1-1
1.2 Com m utation.................................................................................... 1-5
1.3 Performance characteristics................................................ .. 1-12
1.4 Shaft position sensing.................................................................. 1-19

2. MOTOR AND CONTROLLER TYPES


J.R. Hendershot Jr. and TJE Miller
2.1 Introduction...................................................................................... 2-1
2.2 Interior-rotor motors .......................... ......................................... 2-2
2.3 Exterior-rotor motors .................................................................... 2-8
2.4 Pancake or disc-type brushless m o to rs.................................... 2-10
2.5 Slodess m o to rs.............................................................................. 2-12
2.6 Controllersbasic principles ...................................................... 2-14
2.7 The single phaselegthe basic power electronic switch . . . 2-15
2.8 Power semiconductor devices ................................................... 2-18
2.9 Voltage PWM and current regulation............. ........................ 2-21
2.10 Full-bridge circuits for 1- and 2-phase drives.......................... 2-23
2.11 Three-phase full-bridge circuit................................................... 2-27
2.12 Three-phase squarewave control strategies............................ 2-31
2.13 Maximum AC sinewave voltage ................................................ 2-38
2.14 Provision of DC supply voltage ................................................ 2-39
2.15 Controller architecture............................................................... 2-39
2.15.1 Cost issues............................................................................... 2-39
2.15.2 Squaretuave drives...................................................................2-40
2.15.3 Sinewave drives .....................................................................2-46
2.15.3 Unipolar drive........................................................................2-47
2.16 Commutation test circuit............................................................ 2-47

3. BASIC DESIGN CHOICES


J.R. Hendershot Jr.
3.1 Introduction...................................................................................... 3-1
3.2 Interior rotor, exterior rotor, or axial gap configuration? . . 3-3
3.3 Number of phases ......................................................................... 3-3
3.4 Number of stator slots and p o le s....................................................3-4
3.5 Caution with the basic laws of electromagnetism ................ 3-14
vii
Contents
3.6 Simplified motor design ........................................................... S-18
3.6.1 Simple design formulas for the EMF and torque constants . 3-18
3.6.2 Simple calculation of the flu x ..............................................3-20
3.6.3 Basic sizing-rules...................................................................3-23
3.7 Lamination and stator stack design ..................................... . 3-27
3.8 Winding arrangements ............................................................... 3-34
3.8.1 Coil span ............................................... ............................3-34
3.8.2 Coils per p o le ......................................................................3-35
3.8.3 Winding configurationsfractional and integral slot . . . . 3-38
3.8.4 Wire size calculations......................................................... 3-68
3.8.5 Basic winding calculations
3.9 Examples of motor construction .............................................. 3-69

4. MAGNETIC DESIGN
TJE Miller
4.1 Introduction.........................................................................................4-1
4.2 Permanent magnets and magnetic circuits.................................. 4-3
4.3 Approximate calculation of the flux ...................................... 4-10
4.4 Nonlinear calculation of the magnetic circuit.........................4-16
4.5 Armature reaction and demagnetization ............................... 4-18
4-6 Calculation of rotor leakage perm eance................................. 4-23
4.7 Cogging . .......................................................................................4-26
4.9 Retaining can losses .................................................................... 4-28

5. ELECTRICAL DESIGN
TJE Miller
5.1 Introduction.........................................................................................5-1
5.2 Basic windings ....................................................................................5-1
5.2.1 Squarewave motor............................................................... 5-1
5.2.2 Effect of additional coils..................................................... 5-6
5.2.3 Lap windings and concentric w indings............................ 5-9
5.2.4 MuUiple-pole machines........................................................5-11
5.2.5 ConsequentJ>ole windings and m agnets...........................5-12
5.2.6 Computer-aided design of w indings....................................5-13
5.3 Wye and delta connections................................................ 5-14
5.3.1 Wye connection, with 120? trapezoidal phase EMF . . . . 5-14
5.3.2 Delia connection......................................... ..................... 5-21
5.3.3 Flux/pole and magnet utilization......................................5-31

viii
Contents
5.4 Unipolar 3-phase connection..................................................... 5-32
5.5 Two-phase and single-phase connections............................... 5-32
5.6 The EMF constant fcj. ................................................................. 5-35
5.7 The torque constant fcj................................................................ 5-36
5.7.1 Basi* of torque production coenergy............................. 5-36
5.7.2 Torque linearity..................................................................5-38
5.7.3 Demagnetization....................................... .................. 5-38
5.8 Calculating the number of tu rn s .............................................. 5-38
5.9 Winding inductances and armature reaction .......................5-39
5.9.1 Importance of inductance..................................................5-39
5.9.2 Inductance components .................................................... 5-42
5.9.3 Airgap self-inductance of single coil . ................................5-44
5.9.4 Airgap mutual inductance between two c o ils................... 5-46
5.9.5 Examples of inductance calculations............................... 5-48
5.9.6 General case of airgap inductance......................................5-51
5.9.7 Slot-leakage inductance : self and m utual.........................5-55
5.9.8 End-winding inductance.....................................................5-59
5.10 Slotless windings....................................................... ................... 5-61
5.10.1 General ............................................................................... 5-61
5.10.2 Design theory for slotless windings...................................... 5-63

6. SINEWAVE MOTORS
TJE Miller
6.1 Introduction..........................................................................................6-1
6.1.1 The ideal sinewave m otor.................................................. 6-1
6.1.2 Practical motors designed to approximate the sinewave motor 6-2
6.2 Properties of sine-distributed w indings.......................................... 6-4
6.2.1 Conductor and ampere-conductor distributions................. 6-4
6.2.2 Airgap flux produced by sine-distributed w inding............. 6-4
6.2.3 Selfflux-linkage and inductance of sine-distributed witiding 6-6
6.2.4 Mutual inductance between sine-distributedwindings . . . 6-8
6.2.5 Generated E M F .................................................................. 6-9
6.2.6 Torque.................................................................................6-10
6.2.7 Rotating flux and ampereconductors................................6-10
6.2.8 Vector control or "field-oriented" control............................6-11
6.2.9 Synchronous reactance....................................................... 6-12
6.3 Real w indings............................................................................... 6-13
6.3.1 Full-pitch coil.............................................. .......................6-13
6.3.2 Shortjntch co il...................................................................6-14

ix
Co n ten ts
6.3.3 Distribution or spread ....................................................... 6-17
6.3.4 General case ....................................................................... 6-19
6.3.5 Shew ................................................................................... 6-21
6.3.6 Design formulas for practical windings ............................6-21
6.4 Salient-pole m otors...................................................................... 6-23
6.4.1 Calculation of Xd .............................................................6-23
6.4.2 Calculation of X .............................................................. 6-31
6.4.3 Demagnetizing effect of d-axis flux due to Id .................. 6-33
6.4.4 Cross-magnetizing effect of q-axis flux due to I' ............. 6-33
6.4.5 Significance of rotor leakage ............................................ 6-33
6.5 Phasor diagram ............................ ............................................. 6-35
6.5.1 Non-salient-pole machines...................................................6-35
6.5.2 Salient-pole machines.......................................................... 6-36
6.5.3 Operation as a generator ...................................................6-39
6.6 Circle diagram and speed/torque characteristic.................. 6-42
6.6.1 Non-salient-pole motors with Xd ~ ............................... 6-42
6.6.2 Salient-pole motors with Xd ^ X^ .................................... 6-47

7. Kr AND Kg
TJE Miller
7.1 Introduction ................................................. .................................. 7-1
7.2 Squarewave and sinewave m otors................................................ 7-2
7.3 Definition and measurement of fcj- and k ...................................7-3
7.3.1 DC commutator motors........................................................ 7-3
7.3.2 Three-phase squarewave brushless DC motors..................... 7-4
7.3.3 Two-phase squarewave brushless DC motors....................... 7-7
7.3.4 Two-phase sinewave brushless DC m otors.......................... 7-7
7.3.5 Three-phase sinewave brushless DC motors ....................... 7-8
7.3.6 Summary .............................................................................7-10
7.4 Calculation of k j and ............................................................7-10
7.4.1 Squarewave three-phase brushless DC m otors.....................7-11
7.4.2 Two-phase sinewave brushless DC m otors..........................7-13
7.4.3 Three-phase sinewave brushless DC motors .......................7-15
7.5 Example calculation.................................................................... 7-16

8. THE BACK-EMF WAVEFORM


TJE MiUer and R. Rabinovici
8.1 Introduction.....................................................................................8-1
8.2 The flLVmethod ...........................................................................8-4
x
Contents
8.3 Airgap flux-density distribution......................................................8-5
8.4 Skew .................................................................................................... 8-7
8.5 Slotting ................................................................................................8-9
8.6 Calculating back-EMF from tooth flux .................................... 8-11
8.6.1 Single-tooth flux and E M F ................................................... 8-11
8.6.2 Accumulation of tooth fl u x ...................................................8-13
8.6.3 Direct construction of tooth-EAlF waveform........................8-14
8.7 Construction of the phase EMF from the Sj. waveform . . . 8-17
8.8 Development of yoke waveform from (J)T waveform ........... 8-18
8.9 Cogging torque ........................................................................... 8-19

9. CORE LOSSES
TJE Miller and R Rabinovici
9.1 Introduction...................................................................................... 9-1
9.2 Nonsinusoidal Steinmetz equation.................................................9-3
9.3 Core-loss formulas ............................................................................9-3
9.4 Waveform method ............................................................................9-5
9.5 Augmentation of tooth w eight........................................................9-6
9.6 Comparison with test d a ta ............................................................... 9-8

10. ELECTRONIC COMMUTATION OF SQUAREWAVE MOTORS


TJE Miller
10.1 Introduction.................................................................................. 10-1
10.2 Basic principles ........................................................................... 10-2
10.3 Circuit equations - wye ............................................................... 10-7
10.3.1 Commutation................................................................ 10-7
10.3.2 Period A and period B .......... ................................................10-9
10.3.3 Chopping (regulation).........................................................10-10
10.3.4 State-space averaged voltage.f ...............................................10-10
10.3.5 Euler form of voltage equations ..........................................10-11
10.3.6 Initial conditions ................................................................ 10-13
10.4 Circuit equations - d e lta ......................................................... 10-14
10.4.1 Commutation................................................................... 10-14
10.4.2 Period A and period B .........................................................10-17
10.4.3 Chopping (regulation) .........................................................10-18
10.4.4 State-space averaged voltages...............................................10-18
10.4.5 Euler form of voltage equations ..........................................10-19
10.4.6 Initial conditions .................................................................10-20

xi
Co ntents
10.5 Unipolar half-bridge controller ............................................. 10-23
10.5.1 Commutation................................................................... 10-23
10.5.2 Period A and period B ......................................................10-24
10.5.3 Chopping (regulation) .................................................... 10-24
10.5.4 State-space averaged values............................................. 10-25
10.5.6 Initial conditions and final DC values ..........................10-26
10.6 O ver-running............................................................................. 10-27
10.7 Practical examples and comparison with test d a ta ............. 10-28
10.7.1 Comparison of measured and computed waveforms . . . . 10-28
10.7.2 Accurate calculation of no-load speed...............................10-30

11. PERFORMANCE EVALUATION BY TEST


J.R. Hendershot Jr.
11.1 Introduction.................................................................................... 11-1
11.2 Testing of PM brushless m otors.................................................. 11-1
11.2.1 Bach-EMF testing................................................................... 11-2
11.2.2 Resistance and inductance ....................................................11-4
11.2.3 Speed/torque curve and load tests............... ......................... 11-6
11.2.4 Thermal resistance ................................................................ 11.7
11.2.5 Torque linearity......................................................................11-8
11.2.6 Torque ripple..................................................................... 11-10
11.3 Magnetization testing ............................................................... 11-11
11.4 Precision dynamometer ............................................................ 11-12

12. COMPUTER-AIDED DESIGN


TJE MiUer, D.A. Staton and R.P. Deodhar
12.1 The modern design environment ............................................ 12-1
12.2 Basic sizing guidelines................................................................... 12-2
12.3 Computer-aided design with PC-BDC........................................ 12-6
12.4 Finite-element analysis ............................................................. 12-13
12.4.1 Introduction ..................................................................... 12-13
12.4.2 Pre-processing.................................................................... 12-15
12.4.3 Field solution....................................................................12-18
12.4.4 Post-processing.................................................................. 12-19
12.5 Example : armature reaction in brushless DC motor . . . 12-20
12.5.1 Open-circuit flux distribution............................................12-21
12.5.2 Armature reaction field alone........................................... 12-22
12.5.3 Cross-magnetization...........................................................12-25
12.5.4 Demagnetization................................................................12-25
xii
Co ntents

13. EXAMPLES CALCULATED BY HAND


J.R Hendershot Jr.
13.1 Introduction................................................................................... 13-1
13.2 Interior-rotor motor designed from AC induction motor . 13-2
13.3 Exterior-rotor disc drive motor d esig n ................................. 13-16
13.4 Summary .................................................................. 13-26

14. CONTROL SYSTEMS PERFORMANCE


G. Gray
14.1 Introduction.................................................................................. 14-1
14.2 Basic modelling tools for linear control system s................... 14-1
14.2.1 Laplace transforms................................................................14-1
14.2.2 Transfer Junctions ................................................................14-5
14.2.3 Example of a DC or brushless DC motor ............................ 14-5
14.3 Modelling drive com ponents..................................................... 14-7
14.3.1 Brushless PM motor model including inductance............... 14-7
14.3.2 Mechanical and electrical time constants ............................14-8
14.3.3 Transducers..........................................................................14-10
14.3.4 Load effects.......................................................................... 14-10
14.4 Control systems ......................................................................... 14-10
14.4.1 Feedback and closed-loop control......................................... 14-10
14.4.2 Speed controls and servo systems......................................... 14-12
14.4.3 Speed control................................. .....................................14-13
14.4.4 Position control ...................................................................14-13
14.4.5 Torque control.....................................................................14-14
14.4.6 Incremental motion control system s....................................14-14
14.5 Characteristics of closed-loop control systems............. 14-15
14.5.1 Frequency response .................................................................14-15
14.5.2 Bandwidth .......................................................................... 14-18
14.5.3 Step response........................................................................14-18
14.5.4 Stability; gain and phase margins.......................................14-20
14.5.5 Steady-state error................................................................... 14-21
14.5.6 Integral gain compensation.................................................14-22
14.5.7 Root locus.............................................................................14-22
14.5.8 Second-order systems: critical dam ping............................... 14-24
14.6 Control systemsdesign .......................................................... 14-25
14.6.1 Lead/lag compensation ...................................................... 14-26
14.6.2 Pole placement......................................................................14-28

xiii
Co n t e n t s
14.6.3 Robustness ........................................................................14-30
14.7 PID Controllers ..................................................................... .. 14-31
14.7. J Design of a PID controller.................................................14-31
14.7.2 Tuning a PID controller................................................... 14-32
14.7.3 Auto-tuning.......................................................................14-33
14.8 Digital control ........................................................................... 14-34
14.8.1 Discrete system theory........................................................ 14-34
14.8.2 Z-transforms .....................................................................14-36
14.8.3 Z transforms and difference equations........................... 14-38
14.8.4 Stability of discrete systems.................................................14-38
14.8.5 Digital control system design ............................................14-38
14.8.6 Deadbeat controller...........................................................14-39
14.8.7 Digital P ID ........................................................................14-39
14.8.8 PID control example ........................................................ 14-41
14.9 Advanced control techniques ................................................ 1443
14.9.1 Adaptive control ................................................................ 14-43
14.9.2 Optimal control ................................................................14-46
14.9.3 Observers .......................................................................... 14-46

15. COOLING
TJE Miller
15.1 Introduction............. ...................................................................... 15-1
15.2 Heat rem oval................................................................. .. 15-3
15.2.1 Conduction............................................................................. 15-3
15.2.2 Contact resistance...................................................................15-4
15.2.3 Radiation................................................................................15-6
15.2.4 Convection ............................................................................. 15-6
15.2.5 Natural convection................................................................ 15-7
15.2.6 Forced convection ...................................................................15-7
15.2.7 Some rules of thumb for "calibration" ..................................15-8
15.3 Internal temperature distribution ............................................ 15-9
15.3.1 The diffusion equation........................................................... 15-9
15.3.2 Thermal equivalent circuit ..............................................15-10
15.3.3 Current Density ..................................................... 15-12
15.4 Intermittent operation ............................................................ 15-14
15.4.1 Duty-cycle............................................................................. 15-14
15.4.2 Temperature rise during ON-time....................................15-15
15.4.3 Temperature fall during OFF-time.................................... 15-17
15.4.4 Steady-state............................................................. 15-18
15.4.5 Maximum overload fa c to r................................................. 15-18
xiv
CONTENTS
15.4.6 Maximum overload far a single puke ......... .............. .. 15-19
15.4.7 Required cool-down period...................................... .. 15-19
15*4.8 Maximum, on-time,for a given overbadfactor................ 15-19
15.4.9 Maximum duration of single pulse . ,. ............. 15-20
15.4.10 Graphical transient heating curves ........................15-20
15)5 Thermal modelling by computer ............... ........................... 15-22
15.5.1 Computer model of thermal equivalent ckcu.it. . . . . . . 15-22
15.5.2 Determination of equivalent-circuit parameters bytest. . . 15-23

16. MAGNETIC MATERIALS


D.A. Staton
16.1 Introduction .................... .... . ,,.................................................. 16-1
16.2 Permanent magnets ........................................... 16-1
16.2.1 The hysteresis loop and demagnetizationcharacteristic . . . 16-1
16.2.2 Permanent inagnetmaterials . : 3 . ................... 16-4
16-2.3 Temperature effects ............................................................... 16-8
16.2.4 Magnet energy product...........................................................16-9
16.2.5 M agnetization..............................................................16-10
16.2.6 Mechanical properties and handling..........................16-10
16.2.7 The latest trends in magnet technology....................... 16-13
16.3 Soft magnetic iron ................................................................... 16-13
16.3.1 The DC magnetization curve.......................................16-15
16.3.2 Core losses..................................................................... 16-16
16.3.3 Calculation of coefficients for use in core-loss formulas . . 16-17
16.3.4 Work hardening...........................................................16-18
16.3.5 Special steels ................................................................16-19
16-4 Measurement of material characteristics............................. 16-19
16.5 Copper w ire ................................................................. ............. 16-20

xv
INDEX
1. GENERAL INTRODUCTION

1.1 Definitions and types of brushless motor


A brushless motor, as its name suggests, is a motor without brushes,
slip rings or mechanical commutator, such as are required in
conventional DC motors or synchronous AC machines for connection
to the rotor windings, [1 ,2].
Several motors satisfy this basic definition. The commonest is the AC
induction motor, in which current in the rotor windings is produced
by electromagnetic induction. As an AC motor, the induction motor
employs a rotating magnetic field that rotates at a synchronous speed
set by the supply frequency. However, the rotor rotates at a slightly
slower speed because the process of electromagnetic induction
requires relative motion, or slip, between the rotor conductors and the
rotating field. The slip makes the induction motor asynchronous,
meaning that the rotor speed is no longer exactly proportional to the
supply frequency. The induced rotor currents also give rise to I
losses which heat the rotor and decrease the efficiency. Variation of
the rotor resistance with temperature causes the effective torque per
ampere to vary, and this makes the motor more difficult to control,
especially in high-precision motion-control applications. We shall sec
that the brushless permanent-magnet motor overcomes all of these
limitations of the induction motor.
The stepper motor is another common example of a brushless motor.
It has all its windings in the stator, while in most cases the rotor has
permanent magnets and laminated soft iron poles. Torque is
developed by the tendency of the rotor and stator teeth to pull into
alignment according to the sequential energization of the phases
(usually two phases). An advantage of the stepping motor is open-loop
operation: in other words, position control can be achieved without
shaft position feedback. However, in order to achieve stable operation
with adequate holding torque, stepper motors are designed with small
step angles, fine tooth geometry, and a small airgap; these features
tend to increase cost and acoustic noise levels. Stepper motors suffer
from resonances which may occur at multiple speeds in the overall
speed range, and in order to overcome these, inertial dampers or five-
phase configurations are sometimes used. The stepper motor does not

1-1
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

have a fixed torque constant or EMF constant, and it is not amenable to


linear control system design.
The brushless DC motor is essentially configured as a permanent magnet
rotating past a set of current-carrying conductors. In this respect it is
equivalent to an inverted DC commutator motor, in that the magnet
rotates while the conductors remain stationary. In both cases, the current
in the conductors must reverse polarity every time a magnet pole passes
by, in order to ensure that the torque is unidirectional. In the DC
commutator motor, the polarity reversal is performed by the commutator
and brushes. Since the commutator is fixed to the rotor, the switching
instants are automatically synchronized with the alternating polarity of
the magnetic field through which the conductors are passing. In the
brushless DC motor, the polarity reversal is performed by power transistors
which must be switched in synchronism with the rotor position. The
process of commutation is similar in the two machines, and the resulting
performance equations and speed/torque chracteristics are almost
identical.
When the phase currents in the brushless DC motor are of this type, i.e.
DC current which switches polarity in synchronism with the passage of
alternate N and S magnet poles, the motor is said to operate with
squarewave excitation. The back-EMF in this case is usually arranged to
be trapezoidal, and the terms squarewave and trapezoidal are used
interchangeably to refer to the motor and its controller.
There is however, another mode of operation, in which the phase
currents are sinewaves. The back-EMF in this case should ideally be
sinusoidal. Physically the motor and its controller look similar to the
squarewave motor and drive, but there is an important difference. The
sinewave motor operates with a rotatingampere-conductor distribution, similar
to the rotating magnetic field in the induction motor or the AC
synchronous machine. This type of brushless motor is a pure
synchronous AC motor, with fixed excitation from the permanent
magnets. It is more akin to the wound-rotor synchronous machine than
to the DC commutator motor, and for this reason it is often called a
brushless AC motor.
Many brushless permanent-magnet motor products from Japan and
Europe are referred to as AC servo motors rather than brushless DC
machines. Usually when American manufacturers refer to AC machines
1-2
1. G e n e r a l I n t r o d u c t io n
they mean motors which will start and run from the AC line. In this
book, AC implies the use of a rotating ampere-conductor distribution,
which requires a sinusoidal drive. DC implies the use of a commutated
DC drive.
There is some confusion concerning the distinction between brushless
DC and brushless AC motors. This is not an academic question but a very
practical question. The two forms are both manufactured, usually for
different types of application. Their theory, operation, and
manufacturing costs are quite distinct, and in this book the treatment of
the two forms is kept separate.
In Chapters 1-5, the "straightforward approach" to design is developed
in relation to the squarewave motor. The sinewave motor is treated in
detail in Chapters 6 and 7.
STATOR FRAME 2-POLE PERMANENT
MAGNET STATOR

11-SLOT WOUND
ARMATURE (ROTOR)

Fig. 1.1 Cross-section of DC commutator motor. The cxterior-rotor brushless DC motor has
the same cross-section.

1-3
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

The cross-section of a permanent-magnet DC commutator motor is


shown in Fig. 1.1. The stationary magnets are fixed inside the stator
frame, while the rotor carries the rotating winding or armature. Current
is fed to the armature winding by the brushes and commutator. In the
DC machine, the magnetic field is fixed in space. As a result of
commutator action, the current pattern is also fixed in space, even though
the conductors physically rotate. Fig. 1.1 does not show the commutator
or brushes, so it could equally well represent an exterior-rotor brushless
DC motor, with a fixed armature winding on the stator, and rotating
magnets on the outside. This type of brushless DC motor is common in
hard-disk drives for computers. The rotating magnet casing provides a
convenient cylindrical form on which to mount the disk platters, and the
large diameter helps to increase the inertia, which in turn helps to
maintain constant rotational speed.
The cross-section of a typical interior-rotor brushless DC motor is shown
in Fig. 1.2. The magnets are on the rotating rotor. Brushes and
commutator are not necessary because the windings are in the stator and
do not rotate. The smalt rotor diameter reduces the inertia compared to
that of the exterior-rotor motor, and this configuration is common in
servo systems. The stator is similar to that of an AC induction motor. For
low-speed operation it is often sufficient to bond the magnets to the
surface of the rotor hub.
For high-speed operation, the interior-rotor motor may require a
retaining can around the magnets, usually made of thin non-magnetic
stainless steel, or other high-resistivity alloy. The high resistivity minimizes
eddy-current losses (Chapter 4). The retaining can is not shown in Fig.
1.2. Sometimes a kevlar wrap is used instead of a metallic cylinder.
We shall see that there are many variants of brushless motors. For
example, Fig. 1.3 shows three basic configurations. There are axial-gap
disc designs, inside-rotor, outside-rotor and slodess designs, with many
different winding patterns as well as many different pole configurations.
We will review most of the important variants, with insight into the
reasons for the various types and their uses.
The magnets may be in strips, arcs, or discs of various shapes, and they
may or may not be pre-magnetized. Usually, high-energy magnets are
assembled in the pre-magnetized condition, whereas low-energy magnets
can be magnetized after assembly.
1-4
1. G en e r a l In t r o d u c t io n

STATOR (LAMINATED)

Fig. 1.2 Crojj-section of interior-rotor brmhleu DC motor.

1.2 Commutation

The process of commutation is so fundamental to the brushless DC


motor and its control, that it should be understood clearly before
embarking on the design process. In this section the commutation of the
basic squarewave brushless DC motor is described by showing that it is
identical to the commutation of a very simple DC commutator motor.
From this it is shown (in section 1.3) how the two machines have
fundamentally identical performance characteristics.
1-5
D e s ig n o f b r u sh l e s s p e r m a n en t -m a c n e t m o t o r s

STATOR WITH ROTOR WITH


WINDINGS MAGNETS

E U

-ROTOR WITH
c :: id ROTOR WITH
MAGNETS
MAGNETS

STATOR WITH
WINDINGS *1
STATOR WITH
::d WINDINGS------ Px
c : : :d X

UEjd
X

I NNER OUTER AXIAL


ROTOR ROTOR ROTOR

Fig. 1.3 Three basic brashlcss motor configurations

(FROM PM S T A T O R )

Fig. 1.4 Rotor of elementary DC commuUtor motor, rotating in a fixed magnetic field

1-6
1. G e n e r a l I n t r o d u c t io n
Fig. 1.4 shows the rotor of an elementary DC commutator motor,
rotating in a fixed magnetic field. The field is produced by a permanent
magnet, Fig. 1.5. This field is a 2-pole field, because there is only one N
and one S pole in each complete revolution. The axis of the single coil
in Figs. 1.4 and 1.5 is shown at the angle 0 with respect to the reference
axis.

Fig. 1.5 Cross-section of elementary DC commutator motor, showing the position of the
reference axis and the axis of the rotating coil

We need to know first the waveform of the back-EMF in the coil as it


rotates past the magnet. To do this, it is first necessary to find the
waveform of the flux-linkage i|rt, Fig. 1.6a. This is the product of the
number of turns N and the flux ({> passing through the coil. The
subscript 1 denotes the first of three coils or phases.
When 0 = 0 the magnet flux passes either side of the conductors and
does not link the coil. The flux-linkage i|Fj is zero at this position.
Similarly, i|/j = 0 when 0 = 180; and again after one complete revolution
when 0 = 360. In between 0 and 180, the flux-linkage rises to a
maximum positive value at 90, when virtually all the magnet flux passes
symmetrically through the coil. Because of the gap between the N and
S magnets, the coil can rotate a few degrees either side of the 90
position with no change in the flux-linkage. This gives rise to the flat top
in the flux-linkage waveform, Fig. 1.6a.
1-7
D e s ig n o f b r u sh l e ss pe r m a n e n t -m a g n e t m o t o r s

Fig. l.fi Waveforms of flux-linkage. back-EMF, current and torque in an elementary DC


commutator motor or brushless DC motor
1-8
1. G e n e r a l I n t r o d u c t io n
The EMF waveform e1 can now be determined from the ^ waveform by
Faraday's Law, which states that the back-EMF is equal to the rate of
change of the flux-linkage. The instantaneous EMF is therefore
proportional to the slope of the flux-linkage waveform. Mathematically,
rfjr = aijr j b = (11)
dt 36 dt m d9
where ti)ra is the angular velocity. If N is the speed in rev/min, then Ci>m
= 2ti x JV/60 rad/sec. The rate of change of flux-linkage with rotor
position, di|f/30, is obtained from the relevant slopes in Fig. 1.6a, giving
the EMF waveform el shown in Fig. 1.66. In practice, the corners in this
waveform tend to be smoothed out.
The commutator switches the DC supply current tj to the coil with the
same polarity as the EMF, so that the power supplied is gjij. The
waveform of the current ij is shown in Fig. 1.6c. If the angular velocity is
fixed, and if all losses are neglected, the electrical input power is
converted into mechanical power where Tj is the torque
produced by the one coil. The torque waveform is shown in Fig. 1.6d,
and it is clear that unidirectional torque is obtained as a result of the
polarity-reversals of the current in synchronism with those of the back-
EMF.
However, the torque produced by one coil is not constant. There are
periods of zero torque which can be traced back to the flat top in the
flux-linkage waveform. This in turn is due to the physical space between
the N and S magnets.
Fig. 1.7 shows schematically the operation of a DC commutator motor
like that of Figs. 1.4 and 1.5, but with three coils whose axes are
displaced symmetrically at 120 from each other. The coils are connected
together at one point, and their other ends are connected to three
commutator segments, each spanning 120. The torque waveforms
produced by coils 2 and 3 are identical to 7^, but displaced in phase by
120 and 240" respectively, as in Figs. 1.6 and f. The total torque Tis
the sum of Tj + 7g + 7j, as shown in Fig. 1.6g. This torque is constant.

1-9
D e s ig n o f b r u sh l e ss pe r m a n e n t -m a g n e t m o t o r s

Fig. 1.7 Elementary DC motor with 3 commutator segments and 2 brushes


The DC motor with three commutator segments is almost identical to the
three-phase brushless DC squarewave motor. The raifein the commutator
motor are the phases in the brushless motor. The brushes and
commutator are functionally equivalent to the transistor inverter circuit
in Fig. 1.8, and the waveform of Fig. 1.6 apply equally to the brushless motor.

Fig. 1.8 Transistor inverter circuit for use with 3-phase brushless DC motor. The power
transistors perform the switching function of the commutator and brushes, but
they require a separate shaft-position transducer and sensing circuitry (not shown).

1-10
1. G e n er a l I n t r o d u c t io n
Certain important characteristics are the same in both motors. The
current waveform in the phases is a 120 squarewave. Exactly two phases
are conducting at any and every instant. The commutator ensures that
the DC supply current remains constant, as does the torque.
The importance of a constant torque waveform cannot be
overemphasized, especially in servo drives where high precision is
required in both velocity and position control. In machine tools, for
example, the surface finish on machined parts can be adversely affected
by torque variations from the drive motor. Periodic torque variation, or
ripple, is also the cause of vibration which may be extremely troublesome
if it excites a mechanical or structural resonance in the driven equipment
or mountings.
The importance of a constant DC current waveform is that it minimizes
the need for filter capacitors connected across the DC supply, and
generally helps with the levels of harmonics, reducing the filtering
requirements.
At constant speed, the constant torque waveform and the constant DC
supply current waveform represent constant electromechanical energy
conversion according to the equation

E I= T<*m (1.2)
where E is the EMF across 2 phases in series and / is the DC supply
current. The EMF /^connected to the brushes must also remain constant
during each 120 interval, and this can be traced back to the linear rate
of change of flux-linkage of each phase as the magnet rotates. In turn,
this linear rate of change of flux-linkage depends on having a flat-topped
distribution of magnet flux around the stator. More analysis of this is
given in Chapters 5,7,8 and 10.
Equation (1.2) could be said to be the most fundamental equation in
motor theory. It embodies the essential linearity and simplicity of the DC
motor from a control viewpoint, and these are the fundamental features
on which DC servo systems and variable-speed drives have been designed
for many decades. It is very important that the brushless DC motor, in its
ideal form, has exactly the same characteristics.

1-11
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

By contrast, the AC induction motor does not naturally follow equation


( 1 .2), although it can be shown to do so in an indirect sense by means
of mathematical transformations that are the basis of vector control or field-
oriented control The need for these mathematical transformations goes
hand-in-hand with the fact that linear control of the induction motor (as
a servo motor) requires much more complex electronics than the
brushless DC motor, although simple variable-speed induction motor
drives can be made with lower dynamic performance.
A more recent potential competitor to the brushless DC motor is the
switched reluctance motor [3,4]. Although this motor can be controlled
by a simple commutation scheme similar to that of the squarewave
brushless DC motor, this does not result in constant torque or constant
DC supply current, and in order to make the waveforms of these
quantities constant, much more complex controllers are necessary with
current-waveshaping specific to the particular motor.

1.3 Performance characteristics


The best way to gain an understanding of the performance characteristics
of a brushless DC motor is to study the speed vs. torque curve. This
curve represents the steady-state capability of the motor in driving various
types of loads. Its importance stems from the fact that the speed/torque
curve of the motor should be compatible with the speed/torque
characteristic of the load. Certain loads, for example, compressors, hoists,
and conveyors, have a more-or-less constant torque that does not vary
with speed, Fig. 1.9. Others have a torque that increases in proportion to
the speed squared: this is typical of centrifugal pumps, fans and blowers.
Some loads require a constant-torque drive up to a certain base speed, and
a constant power drive at higher speeds. This is typical of traction type
loads, for example, electric vehicles. Washing machines also present this
type of speed/torque characteristic, as do spindle drives on machine
tools.
The most basic function of the speed/torque curve is to ensure that the
motor has enough torque at all speeds from zero speed up to full speed,
to accelerate the load from standstill and maintain full speed without
exceeding any thermal or electrical limits. The thermal and electrical
limits appear as boundaries of regions on the speed/torque curve, as we
shall see.
1-12
1. G en er a l I n t r o d u c t io n

S pEED CENTRIFUGAL TRACTION COMPRESSORS

Fig. 1.9 Typical jpced/torque characteristic] of various loads


The speed/torque curve of a brushless motor can be determined from
a dynamometer test, as described in Chapter 11. In this chapter, we will
develop an equation for the speed/torque curve in its simplest form.
Referring to equation (1.1), the rate of change of flux-linkage can be
seen from Fig. 1.6a to be constant during each 120 conduction interval
on phase 1. It follows that E, the back-EMF from the two conducting
phases in series, can be written
m *EUm (13)

where is a constant called the EMF constant It follows immediately


from equations (1.2) and (1.3) that

T= Jct I, (1-4)
i.e., the torque is proportional to the current, 'rhis proportionality is so
important that the constant of proportionality in equation (1.4) is usually
called the torque constant, Ay. It is clear from this ideal case that kE = k^.
This equality is often overlooked, especially when English or other non-SI
1-13
D e s ig n o f b r u sh l e ss p e r m a n e n t -m a g n e t m o t o r s

units are being used, because then and kj are not numerically equal,
but differ by a constant multiplying factor. Moreover, different test
methods can be used for measuring and fcp and they usually yield
slighdy different values because the magnetic and clectrical conditions
in the motor may not be the same during both tests. (See Chapter 7).
Referring to the circuit of Fig. 1.7, the applied supply voltage Vs is equal
to the sum of the back-EMF E and the resistive volt-drop in the motor
windings, plus the combined volt-drop from two brush/commutator
interfaces:
V3 = E+ RI+ Vh. (J-5)

J? represents the resistance of two coils (phases) in series, and /is the DC
supply current. This equation also applies to the brushless DC motor, if
represents the volt-drop across two conducting transistors in series. In
well-designed systems, is much smaller than the supply voltage Vt, and
for the remainder of this section we will ignore it.
By substituting for E and I, following a little algebraic manipulation the
speed/torque equation can be derived in the following form:

where Q )^ is the no-load speed,

nl -
(1.7)
E

in rad/s and 7 ^ is the locked-rotor torque or stall torque,

Vt (1-8)
^LR = ^T^
in Nm. is the locked-rotor current or stall current, limited only by
the winding resistance.

1-14
l . G e n e r a l in t r o d u c t io n

Fig. 1.10 Speed/torque characteristic of brushless DC motor. The current/torque


characteristic is also shown. The characteristics are plotted for two
voltages.
Equation (1.6) shows that the speed/torque curve is a straight line as
shown in Fig. 1.10. This equation is, in effect, written in normalized
terms: the angular velocity is normalized to the no-load value, while the
torque and current are normalized to their locked-rotor values.
If the motor is operating at no load, the torque is zero and no current
is drawn from the supply. There is no volt-drop in the resistance R and
therefore the motor accelerates until E equals Vs. This occurs, by
definition, at the no-load speed. Equation (1.7) shows that the no-load
speed can be varied by changing the supply voltage: in fact, it is
proportional to the supply voltage. This is precisely the characteristic of
DC motors that made them the mainstay of adjustable speed drives, even
before the days of power electronics. Fig. 1.10 includes a second
speed/torque curve at 1.5 x rated voltage.
When load torque is applied, current is drawn from the supply, resulting
in a volt-drop R im the motor resistance. This volt-drop is possible only
if E falls to the value - RI, and therefore the speed must fall. The drop
in E is proportional to the current, and therefore to the torque, and this
explains why the speed/torque curve at constant voltage is linear.
1-15
D e s ig n o f b r u sh l e ss p e r m a n e n t -m a g n e t m o t o r s

If sufficient load torque is applied, the speed falls to zero and the motor
is then stalled, i.e., in the locked-rotor condition. Then E = 0 and all the
supply voltage is dropped across the motor resistance FL Since R is
usually a very small resistance, the resulting stall current is extremely
large. It is not normally permissible to allow the full locked-rotor current
to flow, even for a short time, because it would either demagnetize the
magnets or destroy the power transistors, or bum the winding
insulation. In fact, normal operation is generally confined to the left-
hand region of Fig. 1.10. Typically, up to 30% of the locked-rotor torque
(and current) may be obtained continuously, and perhaps 50-60% for
very short periods, although these percentages vary widely among
different designs.

Fig. 1.11 Typical continuous and short-time operating regions


Fig. 1.11 shows the typical continuous and short-time operating regions
of a brushless DC motor. The speed/torque curve defines the operating
limits of the motor and its controller. It does not follow that the motor
always operates at a point on the speed/torque curve. In fact, the speed
and torque of most brushless DC motors is continually varying. On
average over a long time period, the operating point must must remain
within the continuous operating range. But excursions into the short-time
range are permitted, and these may be as frequent as necessitated by the
load, provided that the accumulated heating effect does not cause the
1-16
1. G e n er a l I n t r o d u c t io n
motor temperature rise to exceed the short-term rated value. Thermal
calculations are described in Chapter 15. They are extremely important
in rating and selecting brushless DC motors.
Forced cooling, finned housings or both will increase the continuous
power output rating of a brushless DC motor. Cooling has a significant
effect because most of the motor losses producing the heat arise in the
stator, which is the easiest part to cool. In certain aerospace designs,
liquid cooling (fuel or lube oil) is used to cool the stator so that for a
given rating a much smaller framesize can be utilized. There are even
designs which use hollow copper conductors with a cooling liquid passing
through for cooling. Some very high speed brushless motors used for
centrifuges are cooled by refrigeration. These motors frequently operate
in a vacuum to eliminate the friction heating effects of air on the
rotating parts.
The effect of temperature on the magnet characteristics also plays an
important role in the characteristics of the motor. As the motor heats
up, the magnet temperature increases and in most cases this causes a
reduction in the available flux from the magnets. Consequently the
torque constant k j (the torque per ampere) is reduced.
The speed/torque "curves" shown in Figs. 1.10-1.12 are all straight lines,
having been derived from the ideal equation (1.6). In practice, the
speed/torque curve is not straight because of the effects of winding
inductance, which causes distortion at high speed, and because of
magnetic saturation coupled with the demagnetizing effect of the phase
currents at high torque.
Moreover, the ideal analysis presented so far has ignored the effect of all
losses except the 1 losses in the stator windings. Additional losses
include core losses (hysteresis and eddy currents) in the lamination iron,
windage and bearing friction. In addition, there may be eddy currents in
the retaining cylinder if one is fitted; or even in the magnets themselves
in cases where the resistivity is low enough. These additional losses are
caused or increased by variations in magnetic flux-density due to the
stator slotting, or by ripple in the phase currents due to chopping or
pulse-width modulation.
High currents and high temperatures both give rise to concern about
demagnetization. However, in a properly designed brushless DC drive,
1-17
D e s ig n o f b r u sh l e ss pe r m a n e n t -m a g n e t m o t o r s

demagnetization is not a problem because the current is continually


controlled and limited by the controller, and it is a straightforward
matter to detect excessive temperatures in the motor.
It is not obvious from the speed/torque curve how a brushless DC motor
can operate at constant speed, because it appears that the speed falls if
the load torque increases. Remembering that the speed/torque curve is
only a limiting boundary on the operating region, the dotted line in Fig.
1.12 shows constanwpeed operation. As the load torque increases, the
effective supply voltage is increased at the same rate as the R I volt-drop,
so that E remains constant and therefore the speed remains constant.
This can be seen from the fact that the operating point intersects with
a series of spced/torque curves corresponding to the increasing voltage.
The tracking of the RI volt-drop by the adjustable-voltage controller is
achieved by speed feedback control, which is the subject of Chapter 14.
SPEED

1.12 Constant-speed operation of a brushless DC motor, showing that the efFective


supply voltage must vary as the load torque varies.
The perfecdy smooth torque of the ideal brushless DC motor cannot be
obtained in a practical motor, although it can be closely approached.
Torque variations during one revolution arise from imperfect
commutation of the phase currents; from ripple in the current waveform
1-18
1. G e n er a l I n t r o d u c t io n
caused by chopping; and from variations in the reluctance of the
magnetic circuit, due to slotting, as the rotor rotates. This last effect is
sometimes called cogging. It is detectable when the shaft is slowly rotated
by hand. Torque ripple due to imperfect commutation and current
ripple cannot be detected in this way, and it is meaningless to judge the
overall smoothness of a brushless DC motor by turning the rotor on an
exhibition stand! Ways to minimize torque variations are discussed in
later chapters. At high speeds the inertia of the motor and the load
decreases the variation in speed produced by torque ripple of a given
amplitude. At low speeds, a closed velocity loop can virtually eliminate
torque ripple if the feedback gain and bandwidth are sufficiently high.
The audible noise of electrical machines has become an increasingly
important issue. Experience shows that the audible noise of brushless
motors almost always compares favourably with that of any other type of
motor. The reasons for this have to do with the fact that the air gap
between the rotor and the stator is constant and significantly larger than
in induction, stepper, or switched reluctance motors, and that the radial
magnetic forces are not concentrated but are balanced and distributed
over a relatively large pole area.

1.4 Shaft position sensing


The power transistors in the inverter must receive conduction commands
from a system of logic which is synchronized with the rotor position. The
necessary synchronizing signals are usually derived from a shaft position
transducer, and three types are commonly used: the resolver, the optical
encoder, and the IIall-effect transducer.
Fig. 1.13 shows the essence of a resolver system. The resolver provides
very fine resolution in the shaft position signal. Its output is a two-phase
(sine/cosine) signal at the carrierfrequency, modulated sinusoidally by the
rotation of the rotor past the stator. The demodulated output can be
transformed into a pulse train by a resolver-to-digital converter (RDQ,
which is usually a single integrated circuit. Typically, 1000-4000 pulses
per revolution can be obtained. The resolver is an absolute position
transducer because it provides a signal at any position and any speed,
including zero speed. The direction of rotation can be determined from
the relative phasing between the two channels (phases), and analog or
digital speed signals are also readily available.
1-19
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

The resolver is used when more information is needed than just


commutation pulses, especially RPM or tachometer information or a
precise shaft-positioning signal. The resolver is relatively expensive, but
it has the additional advantage of ruggedness and can be used in high-
temperature environments or at high speeds (at least 40,000 rev/m in).
The sensor itself is brushless and may be supplied with or without its own
bearings. The bearingless type (the pancake resolver) consists of a stator
which mounts on the stator frame of the brushless motor.
STATDR
RDTDR

RDTDR STATDR

RE

MODULATING FREQUENCY
<SINE 8. COSINE)
400 Hz INPUT

RDTDR ANGLE

Fig. 1.13 Schematic diagram illustrating the operation of a resolver for shaft
position feedback
The resolver rotor mounts on a shaft extension of the brushless DC
motor at the non-drive end, without couplings. All the necessary
electronic circuits are mounted inside the controller. If the brushless
motor can survive the environment, the resolver can also survive since it
is made up of the same materials of copper and iron, without magnets.
1-20
1. G en er a l I n t r o d u c t io n
Like the resolver, the optical encoder is used when more information is
needed than just commutation pulses. It consists of a set of pairs of
phototransistors and collimated light sources, used in conjunction with
a glass or metal encoder disc. The pattern of slits on the disc defines the
frequency and waveform of the pulse trains which are produced by the
phototransistors.
Encoders can be designed to provide commutation pulses directly,
together with a high-frequency pulse train which may be used to generate
a speed signal.

Absolute

Fig. 1.14 Optical encoder disks: absolute (left) and incremental (right).
Fig. 1.14 shows two types of encoder disks. Commercial incremental
encoders usually have two tracks, A and B, which are in quadrature (out
of phase by one-quarter of a slit-pitch). An index pulse (one slit per
revolution) is also provided as a simple absolute position reference. More
complex encoder discs have special patterns (e.g. Gray Scales) which can
be used to provide absolute position information with very fine resolution
and high accuracy. A typical resolution used in motor drives is 1000
lines/rev.
1-21
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

Optical encoders mount on the motor in a similar way to the resolver,


and versions are available with or without their own bearings and
housings. The direct production of pulses makes the optical encoder
attractive for interfacing to digital commutation circuitry. However, there
are practical limitations. The encoder cannot operate at such high
temperatures as the resolver, and it is not as rugged. Also, the less
expensive incremental encoder (Fig. 1.14) requires an initialization
sequence on start-up, to find the home position. This sequence can be
avoided by using an absolute encoder, but these are more expensive.
Fig. 1.15 shows the Hal]-effect transducer schematically. This is perhaps
the simplest electronic shaft position transducer used for generating
commutation pulses. A Hall switch is a semiconductor switch that opens
or closes when it is placed in a magnetic field higher than a certain
threshold value. It is based on the Hall effect, which is the generation of
an EMF proportional to flux-density when the semiconductor is carrying
current. The EMF, the flux-density, and the current are in mutually
orthogonal directions and the current (usually a few mA) must be
supplied by an external source.
It is common to detect the EMF passing through a threshold value using
signal conditioning circuitry integrated with the Hall sensor, or mounted
very close to it. This provides a TTL-compatible pulse with sharp edges
and high noise immunity, for connection by screened cable to the
controller.
For a three-phase brushless motor, three Hall switches are arranged
spaced at 60 or 120" electrical, and mounted on the stator frame.
Either a separate trigger magnet with the correct pole spacing is
mounted on the shaft in close proximity to the Hall switches, or the Hall
switches can be mounted close enough to the rotor magnets, where they
are energized by leakage flux at the appropriate rotor positions.
A different type of Hall switch is available with a small rare-earth magnet
attached so that the switch conducts current until a steel flux shunt
passes the magnet, causing its flux to divert into the iron path and
turning the Hall switch off. This permits the rotor trigger magnet to be
replaced by a steel trigger shunt.
The various shaft position transducers seem straightforward enough in
themselves, and yet in the marketplace there has been concern with the
1-22
1. G e n er a l I n t r o d u c t io n
use of brushless motors in many applications because of the need for
these devices. On the one hand, it is widely stated that there are no
brushes or commutator to affect reliability. However, these must be
replaced by a shaft position transducer, with additional electronic
circuitry and an interconnecting cable: see Table 1.1. These components
inevitably add to the cost, and may decrease the reliability because they
are relatively fragile and, unless they arc properly protected, they may be
susceptible to damage or maloperation from high temperatures, dust, oil,
vibration and shock, etc., and even from electrical interference.
It is not surprising that there has been much effort in reccnt years to
eliminate the need for the shaft position transducer. At the present
time at least six companies offer control ICs which perform the
commutation without any extra sensors mounted on the motor. This is
called sensorless control. In nearly all cases the principles of the scheme
are the same. For example, in a three phase brushless motor normally
only two of the phases are energized at a time. Since the rotor has
permanent magnets, the back EMF generated in the phase that is not
energized, as well as in the phases which are, can be used for rotor
position information. The shape of the back EMF in the unenergized
phase gives a caricature of the location of the rotor poles and their
magnet polarity. Therefore, these control chips utilize this back EMF as
the analog information which indicates when to commutate the phases.
Several of these integrated circuits are captive to certain markets, and are
not available for general usage.
It is expected that in a few years nearly all brushless motors will use some
sort of sensorless digital control IC. Tachometer or velocity information
is available on these ICs as well as closed-loop speed control and other
features. It should be noted, however, that using back-EMF for shaft
position sensing requires some sort of start-up sequence, because there
is no back-EMF at zero speed. Sequential pulsing of the phases may be
used, but the rotor may start in the wrong direction for a few degrees
before normal control is established. The elimination of this problem is
the subject of development effort, and other proprietary sensorless
schemes are in use which do not depend on back-EMF sensing.
The importance of remote sensing or sensorless control for electronic
commutation for brushless motors cannot be overemphasized and should
always be utilized if possible. The reasons are somewhat obvious: cost,
complexity, extra space, reliability, ability to withstand harsh
1-23
D e s ic n o f b r u sh l e s s pe r m a n e n t -m ag net m o t o r s

environments, and competition from other drive technologies such as AC


induction motors which do not require rotor position feedback. Of
course, servo-quality systems require shaft position feedback, whether
from a transducer or from a "sensorless" position-detection circuit,
regardless of which type of motor is used. However, this is a system
requirement, not a motor requirement.

LOGIC
INPUT v POWER
POWER
AC OR OC

POWER
SUPPLY

Fig. 1.16 Hall-jwitch encoder ichcmatic

1-24
1. G e n e r a l I n t r o d u c t io n

Type of transducer No. of wires in Electronic circuitry


interconnecting cable required in controller
Retoher fi (min) usually RDC (Resolver-lo-digital
screened twisted pairs converter) plus oscillator.
Rugged for high- Typical "S80 + EE PROM &
temperature and/or high control logic. Advantage:
speed applications accurate and valid signals
at switch-on. Good noise
12-16 bit resolution immunity.
Absolute optical encoder 2 supply wires + No. of Power supply for light
bits used, e.g. 8 sources 5-15V depending
Very accurate on optical devices. Hogher
voltage dives better noise
immunity.
Incremental optical 5 (min.) : 2 supply wires + As above. Needs startup
encoder at least 3 others (A,B, procedure to capture
index) index pulse before data is
Typical 1000 lines/rev valid.
Optica] interrupter with 2 + No. of bits As above. No startup
*<pced ring* sequence needed. Poor at
low speed.
Low cost but limited
accuracy and resolution
Hall (witch As above Power supply for Hall
switches: 5 -15V. Signal
conditioning circuit often
mounted In motor.
T a b le 1.1 N o. o f i n t e r c o n n e c t i n g w ire s a n d e l e c t r o n i c s n e e d e d
WITH DIFFERENT TYPES OF SHAFT POSITION TRANSDUCER.
References
1. TJE Miller [1969] Brushless permanent-magnet and reluctance motor drivet, Oxford
University Press. ISBN 0-19-859369-4
2. Kenjo T and Nagamori S [1985] Permanent-magnet and brv.thie.vDC. moton, Oxford
University Press.
3. Miller TJE [1993], Swilchtd-reluctance motor* and their control, Magna Physics/Oxford
University Press, ISBN 1-881855-02-3
4. Miller TJE [1968] Switched reluctance, motor drive*, PC1M Reference Book, Intertec
Communications, Ventura, California

1-25
2. MOTOR AND CONTROLLER TYPES

2,1 Introduction
There are several different configurations of brushless motors which use
rotating permanent magnets and stationary phase coils. The main reason
for so many different variations has to do with the utilization of different
magnet grades in addition to the wide range of applications. For
example, if an application requires rapid acceleration and deceleration
of the load (as in servo systems) then the torque/inertia ratio should be
as high as possible. This indicates the use of an interior-rotor motor with
high-energy magnets.
On the other hand, if an application requires constant speed at medium
to high speed it may make more sense to use an exterior-rotor
configuration with the rotating member on the outside of the wound
stator. This type is sometimes used to drive fans and blowers. Perhaps
the most important application for the exterior-rotor motor is the spindle
motor used in computer fixed-disc drives. This application requires a
very uniform and constant speed, and the high inertia of the exterior
rotor is an advantage in achieving this.
There are other applications such as record players, VCR players, CD
players and floppy disc drives for computers which have a different set
of requirements. These motors rotate at relatively low speed. The
packaging envelope is the most important consideration, and it has been
common to design axial-gap or pancake motors for many of these
applications. In most cases they are slotless, meaning that the magnetic
circuit is closed through a smooth backing plate which may not even be
laminated. Slotless motors are sometimes used in the radial-gap
configuration, particularly in applications where the cogging torque due
to slotting must be eliminated. Slodess motors have reduced core losses
and are suitable for application at extremely high speeds, up to at least
100,000 rpm.

The choice of motor type is the most fundamental design decision,


because of the relatively high cost of magnets, together with issues
related to packaging, magnet retention, and winding.

2-1
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a c n e t m o t o r s

2 .2 In te rio r-ro to r m o to rs
The interior-rotor motor has the closest configuration to that of the
classical AC synchronous machine or the induction motor, although the
production volumes of exterior-rotor motors are much greater. The
stator is similar to that of the three-phase induction motor. Fig. 2.1
shows an example of the interior-rotor configuration. As stated earlier an
advantage of this design is its high torque/inertia ratio. However, it has
two manufacturing disadvantages:
1. Magnet retention must be carefully implemented so that
the rotor does not fly apart.
2. Although exterior stators are easily cooled, they are
expensive to wind without automatic equipment.

ROTOR '

Fig. 2.1 Interior-rotor brushleu permancnt-magnct motor

The rotor shaft must be mounted in bearings. It carries a soft iron yoke
which has a polygonal or circular outside surface on which the magnets
are mounted. The yoke is either machined from low-carbon steel, or
assembled from a stack of laminations (which can be punched from the
2-2
2. M o t o r a n d c o n t r o l l e r types

hole of the stator lamination). Laminations are usually stacked and


pressed on to a light knurl on the shaft to prevent rotation. Magnets are
usually affixed to the yoke surface with care using clean parts, special
adhesives, and appropriate curing. The rotor is often wrapped with
Kevlar yarn or fitted with a tight-fitting steel can to restrain the magnets
at high speeds. Positive magnet retention is essential for high-speed
rotors. The retaining can is usually non-magnetic 300-series stainless steel
or Inconel but magnetic steel cans are sometimes used even though
they increase the rotor leakage. An example is shown in Fig. 2.2c.
Fig. 2.2 shows a selection of interior-rotor designs. Fig. 2.2a shows four
magnets epoxied to a square steel yoke with a 12-tooth stator punching.
The outside diameter of the rotor is ground to the finish dimension after
assembly. Fig. 2.2b utilizes what is known as a bonded ring magnet. These
magnets are available in either bonded ferrite or bonded rare-earth types.
They are less likely to fly apart during rotation, and skewed poles are
easily magnetized on these ring magnets to minimize cogging torque.
Fig. 2.2c shows an 8-pole high-energy magnet rotor using sintered arcs
and a metallic shrink ring or retaining can around the outside of the
magnets. These designs require grinding on the OD of the magnets, the
ID of the ring, and the OD of the ring after assembly.
An interesting variant of the block-magnet or arc-magnet rotor is shown
in Fig. 2.2d. If the polarities of the magnets are like, this design is known
as a consequent-pole rotor. If the polarities of the magnets are unlike, the
motor becomes more of a hybrid between a permanent-magnet motor
and a synchronous reluctance motor, with a high saliency ratio (see
Chapter 6). In both cases only half as many magnets are used, but they
operate with a lower permeance coefficient (see Chapter 4) and must
be thick enough to withstand the demagnetization effect of the airgaps
and the stator current. The particular form of this rotor developed by
Professor Slemon was called the inset-magnet rotor because the magnets
were laid flush in slots let into the rotor surface, giving a rugged
construction that is less dependent on banding for magnet retention [ 1 ].
This design also produces a certain amount of reluctance torque, which
improves the constant-power speed range. In the consequent-pole
configuration the other two poles are soft iron and are integral with the
rotor yoke. The consequent-pole concept can be used with any number
of poles and reduces the cost of the magnets.

2-3
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a c n e t m o t o r s

Fig. 2.2 Example* of interior-rotor brushleu permancnt-magnet rotors

2-4
2. MOTOR AND CONTROLLER TYPES
Fig. 2.2c shows a version of the rotor known as the spoke magnet design or
embedded magnet design. The magnets are magnetized in the
circumferential direction, through their thickness, and their fluxes are
collected and concentrated by soft iron pole-pieces. With 6 or more poles
the flux concentration can be high enough to achieve airgap flux-density
levels, with ferrite magnets, comparable to those which would require a
high-energy magnet in a surface-magnet configuration. The spoke
configuration is becoming more popular because it is a relatively low-cost
fabricated design. The rotor may be larger than that of an equivalent
surface-magnet motor with high-energy magnets, but in many
applications, including many servo systems, a super low inertia is not
necessary. The spoke configuration is especially advantageous in higher-
power machines. Because of the saliency, the most appropriate form of
drive is the sinewave drive (Chapter 6).
The last example shown is Fig* 2.2/which is similar to the first example
except that the magnet arcs are epoxied to a round stack of laminations
so that litde or no grinding is required after assembly unless a retaining
ring is required.

Fig. 2.3 MR! Brkcy-Robinson machine

2-5
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

Another unusual but interesting interior-rotor brushless motor is shown


in Fig. 2.3. The rotor has two steel members which are magnetically
separate. The outer member has two holes through which poles from the
inner member project. Flux leaves the square poles and crosses the
airgap as in a normal synchronous motor, and passes through the stator
yoke in the circumferential direction before returning radially across the
gap to be collected by the outer member of the rotor in the sections
between the holes.
The magnetic potential difference between the outer and inner rotor
members is set up either by permanent magnets or wound field coils
located in the annular space between the two rotors in the regions that
extend beyond the active length. If permanent magnets are used they
may either be stationary or rotate with the rotor. If the magnets are
stationary the necessary clearances introduce additional airgaps. In
principle the rotor of this machine appears to be quite stiff and this
would make it suitable for high-speed applications but for the fact that
the distance between bearings is increased by the rotor extensions.
Moreover, the solid steel is exposed to slot-ripple fluxes that generate
eddy-current losses, and the rotor leakage is inherently high because of
the tortuous flux path.
If the objective is a brushless motor using existing AC induction motor
laminations, windings, and mechanical parts, the use of ferrite magnets
is preferred. The reason for this is that the teeth and yokes of
induction-motor laminations are generally not thick enough to carry the
flux of high-energy rare-earth magnets.
The power density of a ferrite-magnet brushless DC motor may be no
higher than that of the equivalent induction motor, but the efficiency
and full-load power factor will generally be higher, and there is a
significant cost advantage in utilizing the existing tooling and
components of the AC motor. Fig. 2.4 shows the common parts of a
typical induction motor converted into a permanent magnet brushless
motor using ferrite magnets (either sintered or bonded). Motors of this
type manufactured by General Electric (ECM or electronically commutated
motor) are used in high-efficiency domestic heating systems. If a
completely new design had been developed for this application, it is
possible that the amortization of the tooling cost would have been
prohibitive.

2-6
2. M o t o r a n d c o n t r o l l e r types

A L U M IN U M SQ U IR REL
1 PHASE CAGE D I E - C A S T MOTOR
3 PHASE
WINDINGS

2 4 SLOT
STATOR
PUNCH ING
AC IN D U C T IO N MOTOR

Fig- 2.4 Common parts of brushless permanent-magnet and AC induction motors

2-7
D e s ig n o f b r u sh l e s s per m a n en t -m a g n e t m o t o r s

2.3 Exterior-rotor motors


The most cost effective use of ferrite magnets in brushless DC motors
requires a configuration with the rotor outside the stator. Fig. 2.5 shows
the cross-section of a typical motor of this type. The stack of laminations
used for the stator looks very similar to the stack of laminations on a DC
brush-type motor armature.
This type of stator is simple to wind using DC-motor fly-winding
machines. Some designs are produced on winding machines that wind
all three phases simultaneously. The rotor consists of a cup made of soft
iron mounted on the shaft with magnet arcs or a molded or bonded ring
magnet fixed inside the steel rotor cup with epoxy or Loctite. Of
course, balancing is critical because of the large rotating mass. Magnet
retention is provided by the rotor cup on the outside of the magnets. An
advantage of this configuration is the use of a single bearing support of
aluminum or die-cast zinc. Most interior-rotor motors require a bearing
at each end, which results in two bearing end-bells and higher cost.

2-8
2. M o t o r a n d c o n t r o l l e r types

Eg. 2.6 Brush!cm motor integrated in a fan assembly

Fig- 2.6 shows an exterior-rotor brushless motor used in a fan which is


used to cool electronic equipment. These are very efficient, low-cost,
quiet-running fans with the electronics enclosed on the circuit board
within the motor. The assembly is usually mounted with a grille in an
aperture in the casing of a computer or other equipment. They are
sometimes called "muffin" fans and are manufactured in very large
numbers. The smallest are less than 1 in. in diameter. Very long life is
required in many of these designs.
Fig. 2.7 shows another example of an exterior-rotor motor used in
blowers such as the furnace in home heating applications. The
manufacturer of this motor also makes exterior-rotor induction motors
for fan and blower applications. It was natural for them to develop a
similar exterior-rotor brushless machine for improved quietness and
efficiency. The electronic drive is contained within the end of the motor.
This type of motor fits neatly inside the centrifugal blower wheel so that
the passage of air cools the electronics and relieves the motor of the heat
caused by its losses. There is a similar design for "climate-control" blower
motors in automobiles.

2-9
D e s ig n o f b r u sh l e s s p e r m a n en t -m a c n e t m o t o r s

Fig. 2.7 BnuhlcJi blower motor

In general, exterior-rotor brushless motors are used for continuous-


speed applications and the magnet grades are normally the lower-cost
versions of bonded rare-earth, bonded or sintered ferrite grades. Their
popularity continues because of their low cost and ease of manufacture.
2.4 Pancake or disc-type brushless motors
There are many applications for which the packaging of the pancake
brushless motor is extremely convenient. These axtal-gap motors
normally consist of a steel disc rotor with a magnet shaped like a washer
cemented to one side of the disc. The materials used are either sintered
ferrites or bonded rare-earth. They can be easily magnetized with as
many poles as necessary. The stator usually consists o f either printed-
circuit windings or individual wound coils cemented to a printed circuit
board. Fig. 2.8 is a cutaway drawing of a brushless DC pancake motor
used to drive the turntable of a record player. Notice the 6 coils of the
stator mounted to a round plate. The magnet is cemented to a steel
rotor to which the turntable is mounted. To ensure smooth low-speed
performance an extra magnet ring is located on the outer periphery of
the rotor with many poles magnetized on its rim.
2-10
2. MOTOR AND CONTROLLER TYPES

ROTOR YOKE

ROTOR MAGNET

STATOR PHASE
COI LS

HALL SWI TCH

Fig. 2.8 Brushless axial-gap pancake motor for turntable


A pickup coil on the outside of the stator coils generates a tachometer
signal for closed-loop velocity control, creating a very smooth low speed
performance.
There is a very wide variety of configurations for the axial-gap pancake
brushless motor. Fig. 2.9 shows examples.
The main advantages of these motors are their low cost, their flat shape,
and smooth rotation with zero cogging. The air gap between the magnet
and the stator back-iron is very large, and this results in high leakage of
magnetic flux. Although this implies an inefficient magnetic circuit
design, it is not a hindrance for the applications utilizing this motor since
most of these require low torque. However, if axial-gap motors are used
at speeds much above 1000 rev/min, eddy-current losses and heating may
be excessive in the steel backing plate of the stator. There are
possibilities of providing a laminated structure such as a spiral ribbon of
steel or even amorphous iron, wrapped up to form the stator yoke. It is
also possible to eliminate eddy-currents by allowing the backing plate to
rotate with the magnets while the stator coils remain stationary, held by
a non-metallic structure. Such techniques have been adopted for
prototypes or R&D demonstrators of axial-gap motors in larger sizes, but
they can be so complicated as to make the radial-gap motor far more cost
effective in higher-speed, higher-power applications.

2-11
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

Fig. 2.8 Axial-gap brushless motor components. (By kind permission or Sony)

2-5 Slotless motors


Fig. 2.10 shows the configuration of a slotless motor and Fig. 2.11 shows
a finite-element flux plot. The rotor construction is similar to that of the
conventional interior-rotor motor. The stator consists of a stack of
washer-shaped laminations that make a smooth-bore cylinder with a fairly
large airgap that accommodates the windings. The phase coils arc pre
wound before assembly and this permits variations in winding techniques
that are not feasible with slotted stators. For example, helical windings
or Gramme ring windings can be used, with very short end-windings. The
space available for windings is virtually doubled by the absence of the
stator teeth, and this helps to achieve low copper losses, but the flux-
density is reduced because of the large airgap and this machine therefore
naturally has a higher electric loading and a lower magnedc loading than
conventional motors with slotted stators.
The conductors are exposed to the rotating magnet flux and it may be
necessary to use stranded conductors to limit eddy-current losses. Also
the winding must be accurately balanced to prevent circulating currents.
2-12
2. M o t o r a n d c o n t r o l l e r ty pes

STATOR BACK PM ROTOR


I RON

PHASE W I N D I N G S
rig. 2.10 Slodcis brujhlcjj pcrmanent-magnet motor

The main feature of this design is the same as that of the axial-gap
slotless motors discussed in the previous section. Without magnetic teeth
and slots, there is no cogging or reluctance torque. The resulting
performance is very smooth. Torque ripple is still possible by virtue of
the discrete locations of the stator conductors, but winding distributions
with low space-harmonic content can be used to minimize this.
Another version of the slotless motor is a completely ironless slotless
configuration, similar to Fig. 2.10 but the laminated steel stator yoke is
either removed altogether or replaced by a magnetically soft ferrite core
of the same shape. The low saturation flux-density of ferrite may not be
a limitation because the flux-density at its large radius is relatively small.
There are no iron losses with this configuration, and extremely high
speed operation is possible without cogging or heating of the stator core.
An apparent disadvantage of the slotless configuration is the loss of the
heat conduction paths from the conductors to the steel teeth. However,
many slodess prototypes have been built for specialty applications with
direct liquid cooling of the conductors. The inductance of slodess
windings is likely to be low, which makes them suitable for high speed.
2-13
DESIGN OF BRUSHLESS PERMANENT-MACNET m o t o r s

Fig. 2.11 Finite-elcmcnt flux-plot of slotlcu permanent-magnet molor

2.6 Controllersbasic principles


As explained in Chapter 1, the brushless permanent-magnet motor
requires alternating phase currents that may be sinewaves or squarewaves,
depending on whether the motor is designed with sinusoidal or
trapezoidal back-EMF. In both cases the power electronic controllers are
identical, Fig. 1.8, and only the control strategy is different. The control
strategy is the set of rules or algorithms that determine when the power
transistors are switched. The strategy is arranged to give smooth, accurate
control of torque and speed, while limiting the current to a safe value.

2-14
2. M o t o r a n d c o n t r o l l e r types

The brushless permanent-magnet motor is an alternating-current


machine in the sense that the phase currents change polarity every half
cycle as the magnets (which also alternate in polarity) rotate past them.
If the waveform of the phase currents is sinusoidal, the ampere-conductor
distribution rotates continuously at synchronous speed. In order to
achieve sinewave phase currents the DC supply voltage must be switched
on and off at a high frequency under the control of a current regulator
that forces the transistors to switch the correct voltage polarity to the
winding to make the current follow a sinusoidal reference waveform. This
switching of the transistors is called pulse-uiidth modulation.
With squarewaves, the ampere-conductor distribution does not rotate
smoothly but remains fixed in position for 60 electrical degrees, and then
jumps suddenly to a position 60 electrical degrees ahead. The jum p is
effected by a commutation or transfer of current between phases, under
the control of the switching power transistors. During each 60 period
the torque remains constant if the current and the back-EMF both
remain constanthencc the need for a trapezoidal or flat-topped back-
EMF. Maintaining the current constant during each 60 period generally
requires a form of pulse-width modulation or "chopping" under the
control of a current regulator.
Although it may be stating the obvious, it is important to remember that
the power electronic controllers used with brushless permanent magnet
motors are switchmode circuits. This means that the main power
semiconductor switches (transistors and diodes) are operated in such a
way that they are either on or offthe "linear" mode used in linear
power amplifiers is very rarely used. Control is by time-ralio control, also
known as duty-cycle control. The principle of duty-cycle control is that the
power transistors are switched on and off at a high frequency and the
average current is regulated by the relative on-time and off-time.

2.7 The single phaselegthe basic power electronic switch


The basic power electronic switch is the phaseleg, Fig. 2.12, also known as
the half-bridge circuit. The DC supply is shown split into two equal
voltage sources in series, so that the operation of the switch can be
described with reference to the mid-point 0 of the supply. This makes the
operation appear symmetrical in the positive and negative half-cycles.
2-15
D e s ig n o f b r u sh l e s s p e r m a n en t -m a g n e t m o t o r s

Fig. 2.12 The phaseleg or half-bridge circuit, shown with a split-level DC supply.

The single phaseleg is rarely used by itself, but its operadon is considered
here in detail because it forms the basic building block from which full-
bridge and three-phase bridge circuits are derived.
The operation of the phaseleg circuit as a power switch is summarized in
Table 2.1. A single phaseleg has only one output terminal, to which one
end of the load impedance is connected. If there is only one phaseleg,
the other end of the load impedance must be connected to a suitable
terminal so that current can flow in closed loops. In Fig. 2-12 the return
terminal is the mid-point of the supply. If transistor Ql is on and Q2 is
off, the output terminal is short-circuited to the positive of the supply
and the voltage across the load is +Vs/2. Conversely, if Q2 is on and Ql
is off, the output voltage is -Fa/2. Ql and Q2 must never be permitted
to switch on simultaneously because they would short-circuit the supply
and very likely suffer a destructive overcurrent. If Ql and Q2 are both
off, the potential of the output terminal can be either positive or
negative, depending on the condition of the diodes. If Ql was formerly
conducting and switches off, the current commutates to diode D2, and
the inductance of the load keeps the current flowing. The voltage across
the load reverses polarity from +Vs/2 to - V$/2 , and it remains at this
value until the current decays to zero, or until Ql switches on again.

2-16
2. M o t o r AND CONTROLLER TYPES
By switching Q1 on and off at appropriate instants, the average current
can be held within a "hysteresis band of a set-point value. This current
is in the positive or forward direction through the load. Similarly, diode
D1 freewheels the load current when Q2 switches off. Q2 and D1
alternately carry reverse load current When Q2 is on, the load voltage
is ~Vt/2 , and when Q2 is off, it is +Vs/2 for as long as the load current
is freewheeling through Dl.
In a single-phase or two-phase motor, Q1 controls the current for 180
electrical degrees when the motor back-EMF has one polarity, and Q2
controls the current for the other 180 in the cycle when the motor back
EMF has the opposite polarity. There is no overlap between the
conduction periods of transistors Q1 and Q2.
Table 2-1 also shows the direction of current and power flow for the
various conditions of the switches. Since all combinations of voltage and
current polarities are available, and the power flow can reverse, the
switch is said to be a four-quadrant switch. However, reverse power flow
cannot be maintained continuously unless the source is capable of
absorbing it. If the source is a battery, reverse current tends to re-charge
the battery (though not necessarily at the optimum rate). If the source
is a rectifier with a smoothing capacitor, reverse power flow tends to over
charge the capacitor because it is blocked by the rectifier, and the
resulting overvoltage can be dangerous unless auxiliary measures are
taken to absorb the energy, such as crowbar circuits, dynamic braking
circuits, or zener diodes.

Q2 Dl D2 4d Power
1 0 0 0 +V./2 + +
0 1 0 0 - V s - +

0 0 0 1 -V./2 + -
0 0 1 0 + V /2 - -
0 0 0 0 Floating 0 0
1 1 0 0 Shoot-through

T a b le 2.1 O p e r a t io n o f s in g l e p h a s e l e g c ir c u it

2-17
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

2.8 Power semiconductor devices


By operating the power transistors only in the on-state or the off-state,
the conduction losses in the power transistors are minimized. These
losses depend on the product of device voltage and device current. When
a power transistor is on, the device voltage is close to zero, and when it
is off, the device current is close to zero. Consequently the voltage-
current product is close to zero at all times except during the transition
from on to off, or from off to on. The transient losses during the
switching transitions are called switching losses. They are proportional to
the switching frequency and also to the transition times during turn-on
and turn-off. The best power transistors have very rapid turn-on and turn
off times, of the order of nanoseconds or microseconds, and device
manufacturers are continually striving to make them even faster.
The ideal power semiconductor switch has a number of desirable
properties:
1. Zero on-state forward voltage dropnot only to minimize conduction
losses, but also to minimize the "wastage" of supply volts which could
otherwise be used to force current into the motor. This "wastage" of
supply volts is critical in low-voltage systems such as automotive
applications that run from 12V supplies. Some electronic equipment runs
from supply voltages even lower than 6V.
2. Zero leakage current m the offstateto minimize losses due to the product
of leakage current and device voltage. When a power transistor is "off1
the voltage across it may be high, so even a small leakage current can
generate significant losses.
3. High forward-blocking capabilitythe forward blocking capability should
be higher than the supply voltage by a safety margin that depends on the
types of transients that are expected. Typically the "rated" voltage of
devices is at least 30% above the supply voltage, but larger margins may
be necessary to accommodate adverse conditions such as overvoltage on
the supply, or high temperature. Reverse blocking capability is generally
only a fraction o f forward blocking capability (except in diodes and some
power MOSFETs) but in most of the bridge-type circuits used with motor
drives, reverse blocking is not needed because the transistors are
protected from reverse voltage by appropriately-connected freewheel

2-18
2. M o t o r AND CONTROLLER TYPES
diodes. Of course the diodes must have adequate reverse-blocking
capability comparable with the forward-blocking capability of the
transistors, since they are generally connected in anti-parallel with the
transistors.
4. High dv/dt capabilitytransistors should be as near as possible immune
from spurious turn-on caused by the induction of gate current caused by
high dv/dt across the main power terminals. Modern power transistors
are generally MOS-gated, with capacidve input impedance at the gate,
and therefore are inherently suscepdble to spurious turn-on if the gate
is subject to a high dv/dt, which may be coupled via the Miller
capacitance between collector and gate (in an IGBT) or between source
and gate (in a power MOSFET). High dv/dt immunity in the device itself
is obviously desirable, but the safest policy is to drive the gate from a low-
impedance source/sink (such as a pair of driver transistors connected in
totem-pole fashion).
5. High di/dt capabilityM power devices have a maximum rate of change
of current that can be tolerated without current-crowding effects or
second breakdown.
6. High speed of switchingin transistors, the turn-on and turn-off times
need to be as fast as possible to minimize switching losses, although there
is no point in paying a premium for very fast transistors if the switching
losses are insignifcant compared with the conduction losses. In diodes,
the transition from off to on needs to be as fast as possible because the
commutation of inductive current from a transistor branch to a diode
branch is the main means for protecting against destructive transient
voltages. Diodes should also have good reverse-recovery characteristics
(see Chapter 10).
The most startling aspect of the development of brushless motor drives
is the sustained rate of improvement in power electronic switches and in
integrated circuits for digital control. The first brushless motors arrived
on the scene in the early 1960s. They used Alnico magnets, and their
commutation sensors comprised filament lamps with planar silicon photo
transistors. The first power electronic switching device to be developed
and widely applied in motor drives was the SCR in the 1960s. Because
the SCR cannot be turned off by a gate signal, it lent itself to naturally-
commutated converters such as phase-controlled rectifiers used with DC

2-19
D e s ig n o f b r u sh l e ss p e r m a n e n t -m a g n e t m o t o r s

commutator motors, or load-commutated inverters used with synchronous


motors. Although forced-commutated inverters were developed and
widely used with AC induction motors, they required bulky commutation
circuits and the maximum switching frequency was limited to a few
hundred Hz.
The brushless permanent-magnet motor requires current regulation that
is possible only with power semiconductor devices that can turn off as
well as turn on, preferably at frequencies of 10 or 20 kHz, and therefore
the development of brushless motor drives did not really take off until
the introduction and rapid development of the bipolarjunction transistor
in the 1970s. Since that time the BJT has been steadily overhauled by
MOS-gated power devices because of the great simplification in gale drive
circuitry. The most important of these devices are the power MOSFET
and the insulated-gate bipolar transistor (IGBT). The MOSFET is the
fastest-switching power device and permits the highest switching
frequency. For low-voltage systems it can be designed with a very low on-
state resistance, of the order of 0.006 Ohms or less in a 60V rated device
(although this increases with increasing temperature). Although
MOSFETs are available up to several hundred volts and tens of amperes,
the IGBT is more suitable for high-power applications. It is available in
voltage ratings up to 1700V (with much higher values expected) and with
current ratings of 500A or more. For the future, MOS-controlled
thyristors (MCTs) are being developed for higher-voltage applications
especially where high junction temperatures are likely.
At the time of writing the technical capabilities of power electronic
devices are well matched to the requirements of brushless drives, at least
up to several tens of kW. The chief limitation to the wider deployment
of brushless drives is the cost of the electronic controller including its
sensors (current sensors and rotor position sensors), especially when
compared with very low-cost configurations such as the triac-controlled
AC commutator motors used in washing machines. However, for
comparable performance the brushless permanent-magnet motor is at no
disadvantage with respect to controller cost compared with the induction
motor or the switched reulctance motor, and in many applications
(particularly those where energy saving is important) it has the advantage
of higher power factor which translates into a lower volt-ampere rating
in the controller. This is most important in drives of larger sizes,
although induction and SR motors improve at higher power levels and
of course they have no magnet cost.
2-20
2. M o t o r a n d c o n t r o l l e r types

Load current

Eg. 2.13 Voltage PWM

2.9 Voltage PWM and current regulation


Voltage pulse-width-modulation (PWM) is shown in Fig. 2.13. The power
transistors are switched on and off, usually at a fixed switching frequency
fh. The switching period is Ts = \ / f t. During the on-time *qN, the voltage
applied to the load is V5. During the off-time /Qjy, the voltage applied to
the load is either zero or - If it is zero, the "chopping" of the voltage
is called soft chopping. If it is - VE, the chopping is called hard chopping. Soft
chopping requires a "zero-volt loop" in which the load is short-circuited,
usually through one transistor and one diode. Hard chopping requires
that the voltage across the load be reversed, usually through two
freewheel diodes. The duty cycle d is defined as
d = = ___'N . ( 2. 1 )
T& 'o n + 'o f f
If the supply is fixed and the duty-cycle is also fixed, it is not difficult to
show that the average voltage applied to the load is
* W l = 4 * Vs- (2-2 >

2-21
D e s ig n o f b r u sh l e s s p e r m a n en t -m a g n e t m o t o r s

The duty-cycle can be modulated or variedfor example, sinusoidally, so


that the load current is sinusoidal. If the modulation is at a certain
frequency /[, it is desirable that the switching frequency ft be much
highergenerally ten times higher, so that the inductance in the load will
reduce the switching-frequency ripple in the load current. For example,
if the fundamental frequency of current supplied to a brushless motor is
200 Hz, the switching frequency should be at least 2 kHz. At lower power
levels it is feasible to switch at much higher frequencies, above 20 kHz,
so that very smooth current waveforms are obtained. Switching
frequencies at this level are also inaudible, which is an important
advantage.

Load current limited to iH i

\A A A A A A 7 \7 ~ y Hysteresis
band

Fig. 2.14 Current regulation

Current regulation in its simplest form is shown in Fig. 2.14. The


regulated current is passed through a current transducer which should
have sufficient bandwidth to follow rapid variations in current. The
current feedback signal is compared with a reference or set-point current
in a comparator which has a small amount of "hysteresis". When the load
current exceeds the set-point value the driving transistor is switched
off and the current falls, owing to the motor back-EMF and the resistance
drop in the circuit. When the current falls below the level the driving
transistor is switched back on again. The current follows the hysteresis
band - ij 0), with a ripple frequency and amplitude that depend on
the load inductance, the circuit voltages, and the bandwidth of the
current-regulating circuitry.

2-22
2. MOTOR AND CONTROLLER TYPES
The switching of the power transistor is not instantaneous when the
current crosses the threshholds and and therefore the current
waveform can leak outside the hysteresis band. By how much depends on
the response rate of the controlling circuit Moreover, the switching
instants do not occur at a fixed frequency and if it is necessary to filter
the switching frequency to protect neighbouring circuits, the job may be
made more difficult because of the frequency variation.
The two switching control strategies, voltage-PWM and "current
regulation", are only the simplest examples of many possibilities. Much
more sophisticated implementations have been developed, using such
techniques as purely digital regulators, pre-programmed switching intants
to minimize the harmonic content in the current waveforms, and
regulators whose switching is defined and implemented in a rotating
reference frame.

2.10 Full-bridge circuits for 1-, and 2-phase drives


Fig. 5.16 shows the jutt-bridge circuit feeding a single-phase load, for
example, one phase of a brushless motor. The switchmode operation of
this circuit is summarized in Table 2.2. There is no need to split the
supply voltage into two equal halves. When Q1 and Q2 are on, the
voltage across the load is the full supply voltage Vt. When Q3 and Q4 are
on, the load voltage is -V t. Compared with the half-bridge circuit of Fig.
2.12, the voltage available to drive the load is doubled, even though the
DC source voltage is the same. For the same power, the motor can be
wound with half the number of turns and the current is halved.
The first four rows show the control of transistors Q1 and Q2 to control
positive or forward current. Diode D4 freewheels the load current when
Q.1 is off, and D3 freewheels the load current when Q2 is off. In both
cases with one diode and one transistor conducting, the load is in a
"zero-volt loop" and the freewheeling load current decays under the
influence o f the motor back-EMF and the resistance volt-drops in the
circuit. These loops permit the implementation of soft chopping, which
minimizes the current ripple for a given chopping frequency, or
alternatively minimizes the chopping frequency required to limit the
current ripple to a given level. W hen both transistors are off, the load
current freewheels through D3 and D4 and these diodes connect the
negative of the supply voltage to the load terminals. During the
2-23
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

conduction period for forward current the transistors Q3 and Q4 and the
diodes D1 and D2 are idle.
The second group of four rows shows the identical control technique
applied to transistors Q3 and Q4 for negative or reverse current. During
the conduction period for reverse current the transistors Q l and Q2 and
the diodes D3 and D4 are idle.

Ql Q2 Q3 Q4 D1 D2 D3 D4 ^load Power
l 1 0 0 + +

0 1 0 1 0 + 0
l 0 0 0 1 0 0 + 0
0 0 1 I + -

1 1 0 0 - +
0 1 0 1 0 - 0
0 1 0 1 0 0 0 - 0

0 0 1 1 +V, -

T able 2.2 O p e r a t io n o f f u l l -b r id g e c ir c u it

Fig. 2.15 shows the waveforms of applied voltage and current for a full-
bridge circuit controlling a single-phase load with 180 conduction for
forward and reverse currents. The diagram shows the on-periods of each
transistor, and the chopping waveforms for forward conduction. The
diagram shows one transistor (Ql) chopping for the whole 180 interval,
while Q2 remains on during the whole of this interval. In this case Ql is
called the chopping transistor and Q2 is called the commutating transistor.
The chopping duty can be assigned to one transistor or it can be shared
between the two. For example, in the next half-cycle of forward current,
Ql could remain on while Q2 chops. The load voltage would be the
same as in the first half-cycle, but the thermal duty experienced by both
transistors would be equalized because their mean and RMS currents
would be equalized.

2-24
2. M o t o r a n d c o n t r o l l e r ty pes

Una Current |
Cl
4
o 90 80 80 120 150 180 210 240 270 300 990 MO Etecdagraaa

Phw* EMF

eph llne
*ph" ^llne

Q1 Q3 Q1

Q4 Q2 Q4

Q1
duty-cycle d JL T U T JU L
Chopping transistor
D4
duty-cycle 1-d n r ^ ir
Chopping c iode
Q2

Commutating transistor
D3

Commutating diode

fig. 2.15 Voltage and current waveform* for fuU-bridge circuit controlling a single
phase load with 180 conduction periods for forward and reverse current

2-25
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

The waveforms in Fig. 2.15 can be used as the basis for calculating the
peak, mean, and RMS currents in the load, in the chopping transistor,
and in the commutating transistor and their respective diodes. Table 2.3
summarizes the result for the case where one transistor does all the
chopping, and Table 2.4 for the case where the chopping is shared
between the two transistors of each phaseleg on alternate cycles. All
these currents are expressed per-unit of the set-point current 7sp which
is the flat-topped value of the line current in Fig. 2.15.
Peak Mean RMS
Line 1 1 1
Chopping transistor 1 d/2
Chopping diode 1 (1 -d)/2 /[ (l- d ) /2 )
Commutating transistor 1 1/2 1 //2
Commutating diode 1 0 0

T able 2.3 P ea k , m ea n a n d rm s c u r r e n t s sin g l e p h a s e f u l l -b r id g e


c ir c u it , n o t t h er m a ll y e q u a l iz e d . S q u a r ew a v e d r iv e .
(d = duty^ycle).
Peak Mean RMS
Line 1 1 1
Transistor 1 (1+ 4/4 /[l+ ifl/2
Diode 1 (l-rf)/4 /[l-rfJ /2

T a b le 2.4 P eak , m e a n a n d r m s c u r r e n t s s in g l e ph a se f u l l - b r id g e
c ir c u it , th e r m a l l y e q u a l iz e d . S q u a r ew a v e d r iv e .
(d = duty-cycle).
In two-phase drives the operation is similar, but each transistor conducts
for only 90 instead of 180. The waveforms are shown in Fig. 5.17. With
only one transistor chopping in each phaseleg, the peak, mean and RMS
currents are summarized in Table 2.5. If the transistors chop on alternate
cycles, for thermal equalization, the peak, mean, and RMS currents are
as summarized in Table 2.6.

2-26
2. M o t o r a n d c o n t r o l l e r types

Peak Mean RMS


Line 1 1 1 //2
Chopping transistor 1 rf/4 /[rf}/2
Chopping diode 1 (1 -d)/4 /[ (l- r f )] /2
Commutating transistor 1 1/4 1/2
Commutating diode 1 0 0

T able 2.5 P eak , m ean and rm s cu rren ts w o -ph ase fu ll -bridge


CIRCUIT, NOT THERMALLY EQUALIZED. SQUAREWAVE DRIVE.
(d = duty-cycle).
Peak Mean RMS
Line 1 1 1
Transistor 1 (l+rf)/6 /[l+ < i]/2 /2
Diode 1 <l-rf)/B /[ l- ( f l/2 /2

T a b le 2.6 P eak , m ea n a n d rm s c u r r e n t s t w o -p h a s e f u l l -b r id g e
CIRCUIT, THERMALLY EQUALIZED. SQUAREWAVE DRIVE.
{d = duty-cycle).

2.11 Three-phase full-bridge circuit


The three-phase full-bridge circuit is shown in the upper diagram in Fig.
2.16 with wye-connected motor phase windings 1,2 and 3; and in the
lower diagram with delta-connected phase windings. In both cases the
power electronic circuit and the switching sequence of the transistors
remain the same. The lines are labelled A,B and C.
Squarewave operation is shown in the upper part of Fig. 2.17. In
squarewave operation DC current is fed from the supply to the motor via
two lines for an interval of 60. During this interval the third line carries
no current and is idle. At the end of each 60 period the current
commutates from one of the conducting lines into the idle line. There
are normally 2 transistors conducting at any time: an upper and a lower.

2-27
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

Fig. 2.16 Three-phase bridge circuit for sinewave and squarewave drives, (a) wye
connected motor {!>} delta-connected motor

2-28
2. MOTOR AND CONTROLLER TYPES
-30 0 30 60 90 120 150 180 210 240 270 300 330
!SP alec*
1A

*B

*C

I
Q5 Q1 03 Q5
G6 Q2 Q4
ABC A BC A BC A BC ABC ABC
W W Lit LU tu tu

Fig. 2.17 Line current waveforms for three-phase squarewave (upper) and sinewave
(lower) drive*, including the states of the transistors and current paths.

2-29
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

Fig. 2.17 shows the states of the power transistors Q1-Q6, and the arrow
diagrams indicate which lines are conducting in which direction. The
angles at the top of Fig. 2.17 represent rotor position in electrical
degrees.
Sinewave operation is shown in the lower part of Fig. 2.17. In sinewave
operation, the line currents are essentially sinusoidal although they have
a ripple component due to the chopping of the transistors. Normally all
three currents are non-zero (except when they pass through zero).
Consequently there are normally three transistors conducting at any time:
either one upper and two lower, or two upper and one lower. Fig. 2.17
shows the states of the power transistors Ql -Q 6, and the arrow diagrams
indicate which lines are conducting in which direction.
In both the squarewave and the sinewave drives it appears from the ideal
waveforms in Fig. 2.17 that there is really only one DC current which is
switched or commutated among the phases. This seems to imply that the
current could be measured with only one current sensor in the DC
supply, and regulated by chopping only one transistor. However, the
operation of the circuit is complicated by the action of the freewheel
diodes and the motor back-EMF. Because of the diodes, the three phase
currents are not necessarily "observable" to a current sensor in the DC
supply. For full control of the current at all times, it is usual to measure
the line currents directly by means of current sensors in the lines.
However, in a three-wire connection iA + ijj + ^ = 0 so that only two
currents need be measured: the third can be determined from the sum
of the other two. This saves one current sensor.
When the control strategy is voltage PWM in the sense described in
connection with Fig. 2.13, it is common practice to use only a single
current sensor in the DC supply, where it can be used not only for
overcurrent protection but also in a torque control loop.
The use of current sensors in the lines does not necessarily guarantee the
detection of overcurrents in all branches of the circuit, since this
depends on the association of the current-feedback signal with the gate
control of the transistor that is chopping at the time.
A full analysis of the transistor and diode currents for 1-, 2- and 3-phase
circuits is given in Chapter 10.

2-30
2. MOTOR AND CONTROLLER TYPES

"build" loop freewheeling loop

Fig. 2.18 3-phase bridge circuit showing conducting loops just after Q5 has turned
off and Q1 has turned on. This is the start of the 60 "base interval".

2.12 Three-phase squarewave control strategies


If current is being regulated, it is the line current. If voltage is being
regulated, it is the line-line voltage. In most respects the control strategy
is identical for both wye and delta connections, and the ideal line current
waveforms are the same in both cases, Fig. 2.17.
2-31
D e s ig n o f b r u sh l e s s p e r m a n en t -m a c n e t m o t o r s

Six different control strategies will be described. Four of these regulate


the line currents. The other two regulate the line-line voltage. In all
cases, however, the commutation sequence is the same as in the upper
diagram in Fig. 2.17.
The current waveforms are divided into 60 intervals, which are
essentially identical except for the commutation or re-distribution that
takes place at the end of each interval. Any interval can be constructed
from one 60 interval chosen as the "base" interval, by "commutating"
the currents correctly according to the rotor position. The base interval
is chosen as the 60 interval following the turn-off of Q5. This is also the
start of the 120 conduction period for Ql, and it is the instant when tA
takes over the DC conduction from i^.
Fig. 2.18 shows the currents flowing immediately after the start of the
base interval. The current in line C was flowing in Q5, and is now
freewheeling through D2. The current in line A is building up from zero.
The base interval is divided into two periods, A and B. Period A is the
freewheeling period of D2, and ends when the current in D2 reaches
zero. Period B is the remainder of the 60 base interval. Normally,
Period A is much shorter than Period B, except at high speeds or when
the motor has excessive inductance. (See Fig. 10.7).
The control of current or voltage is by one of the following six options.
In all cases, only one transistor is used for chopping at any time. This
may be an upper or a lower transistor.

C 120 Q l The "C" means that line current is the regulated parameter.
Ql is the control transistor (chopping transistor). Note that
it is the "incoming" transistor. Throughout the base interval
it is paired with Q6, which is the "old" or "outgoing"
transistor. Q6 remains on for the whole 60, and turns off
at the end (i.e. at the rotor position 90 in Fig. 2.17). In
this strategy, only the upper transistors are used for
chopping. Therefore, Ql continues to chop for 120. Q6
is used only for commutation, and does not do any
chopping. Q3 and Q5 are therefore equivalent to Q l, and
have the same peak, mean, and RMS currents as Q l. Q4
and Q2 are equivalent to Q 6.
2-32
2. MOTOR AND CONTROLLER TYPES
The current sensor controlling Q1 is assumed to be in line
A or in the DC supply line. This ensures that iA will not
exceed the set-point current /sp. However, this sensor does
not detect the current in line B, which is the sum of the
rising current iA and the freewheeling current from line C.
Therefore, i0 can exceed 7sp, and indeed "spikes" of excess
current can be seen in the line current waveforms with this
strategy, as illustrated in the smaller diagram in Fig. 2.18.
Note also that the freewheeling current in Fig. 2.18 is in a
"zero-volt loop", and therefore its decay rate depends on
the total back-EMF in that loop. Under certain condidons
this decay rate may be too slow, so that D 2 never turns off
during the base interval.
The ideal current waveform in Q1 comprises the positive
half-cycles of the current waveform shown at the top of
Fig. 2.17, with chopping continuing for 120. This
waveform is reproduced in Fig. 2.19.
C 60 Q1 This strategy is similar to the previous one, but with Q1
(the incoming transistor) chopping for only 60 to control
the current in line A. The ideal current waveform is shown
in Fig. 2.19, together with the chopping diode current in
D4.
All six transistors do identical duty. Each one conducts for
a total of 120, but it is chopping only for the first 60 of
its conduction period. In the second 60, it remains on,
and the chopping is continued by the "complementary"
transistor, which is always the next one in the commutation
sequence (i.e., Q l,2,3,4,5,6).
V 120 Q1 The "V" means that line-line voltage is the regulated
parameter. Q 1 is the control transistor (chopping
transistor). Note that it is the "incoming" transistor.
Throughout the base segment it is paired with Q 6, which
is the "outgoing" transistor. Q 6 remains on for the whole
60. In this strategy, only the upper transistors are used for
chopping. Therefore, Q1 continues to chop for 120. Q 6
is used only for commutation, and does not do any

2-33
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

chopping. Q3 and Q5 are therefore equivalent to Ql, and


have the same peak, mean, and RMS currents as Q l. Q4
and Q2 are equivalent to Qj6.
No current sensor is used for instantaneous current
regulation. Therefore the duty-cycle d and chopping
frequency must be adjusted to achieve the desired current.
The ideal current waveform in Ql is shown in Fig. 2.19,
with chopping continuing for 120. The corresponding
diode current in D4 is also shown. The waveform is similar
to that of the C 120 Q l strategy, except that the chopped
current waveform is not clipped at 7sp but varies depending
on the back-EMF, inductance, speed etc. In general, the
chopping duty-cycle d should be increased approximately
in proportion to motor speed to keep the average applied
voltage above the back-EMF.
C 120 Q6 The "C" means that line current is the regulated parameter.
Q 6 is the control transistor (chopping transistor). Note that
it is the "outgoing" transistor. Throughout the base
segment it is paired with Q l, which is the "incoming"
transistor. Ql remains on for the whole 60. In this
strategy, only the lower transistors are used for chopping.
Therefore, Q6 continues to chop for 120. Q l is used only
for commutation, and does not do any chopping. Q3 and
Q5 are therefore equivalent to Ql, and have the same
peak, mean, and RMS currents as Q l. Q4 and Q2 are
equivalent to Q 6.
The current sensor controlling Q6 is assumed to be in line
B. This ensures that iB will not exceed the set-point current
7SP. This constitutes the main difference between this
strategy and the C 120 Q l strategy. It also means that the
rate of rise of current in line A may be limited (by
chopping Q 6) to maintain iB < /sp. (With the C 120 Q l
strategy, there is no limit on the rate of rise of iA before it
reaches 7sp. When Q 6 switches off during chopping, the
freewheeling current in line C returns to the supply via D3
(and D2). During these off-periods the supply voltage will

2-34
2. M OTOR AND CONTROLLER TVTES
lend to extinguish the freewheeling current quickly.
The ideal current waveform in Q6 is shown in Fig. 2.19,
with chopping continuing for 120. The corresponding
diode current in D3 is also shown.
C 60 Q 6 This strategy is similar to the previous one, but with Q6
(the "outgoing" transistor) chopping for only 60 to
control the current in line A. The ideal current waveform
is shown in Fig. 2.19, together with the chopping diode
current in D3.
All six transistors do identical duty. Each one conducts for
a total of 120, but it is chopping only for the second 60
of its conduction period. In the first 60, it remains on,
and the chopping is continued by the "complementary1
transistor, which is always the previous one in the
commutation sequence (i.e., Q l,2,3,4,5,6).
V 120 Q 6 This is similar to the V 120 Q l strategy, except that the
"outgoing" transistor is used for chopping instead of the
"incoming" transistor.

Selection of Appropriate Switch Control Strategy


If a current-regulating loop is used to control the instantaneous current,
a C strategy should be selected. If voltage-PWM is used, a V strategy is
appropriate.
If all transistors do identical duty, a 60 strategy should be used. If only
the uppers (or only the lowers) are used for chopping, a 120 strategy
should be used.
If the current sensor is in the DC supply line, a Q l strategy should be
used. If the current sensor is in each individual line, a Q 6 strategy should
be used.
The peak, mean and RMS currents in the transistors with three-phase
squarewave drive are summarized in Tables 2.7 and 2.8.

2-35
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a c n e t m o t o r s

tsga
Current
chopping
m m
Chopping
diode
VA7
Voltage PWM
continuously
iON . i
0 30 60 90 120150 180 210 240 270 300 330 360 elec
------------------------------ Lina
la Currant

^_1
D4 |-
;: ;| C 120 Q1

-01______________________ 1
D4 I C 60Q1

_Q1 V/////Z77777A
D4 r ----. 'i V 120Q1

06 C 120 06
L ..: ' : i 1 D3

06 C 60 06

03
YZZZZZZZZZZZA 06 V 120 06
r D3

05 01 03 05
06 -92. 04
Fig. 2.19 Effect of switch control strategy on mean and RMS transistor and diode
currents in three-phase squarewave drives.

2-36
2. M o t o r and c o n t r o l l e r ty p e s

WYE mad DELTA


Peak Mean RMS
Line 1 2/3 4 2 /3 ]
Chop, transistor 1 rf/8 Ad/%]
Chop, diode 1 (l-rf)/3 '/[(l-rf)/3]
Comm, transistor 1 1/3 A U i]
Comm, diode 1 0 0

DELTA
I Peak Mean RMS
Phase 2/3 4/9 4 2 1 /3

T able 2.7 P eak , m ean an d RMS c u rren ts in squarew ave t h r e e -ph a se


c o n t r o l l e r C 120 Q l , C 120 Q6, V 120 Q l, AND V 120s
Q6 CONTROLS. (d = duty-cycle).

WYE and DELTA


Peak Mean RMS
Chop. & comm, 1 (l+rf)/fi v'[(l + <i)/6]
transistor
Chop, & comm, 1 (l-d)/fi v '[(M /6 ]
diode

T able 2.8 P eak , m ean and RMS c u r r en ts in squarew ave t h r e e -ph a se


c o n t r o l l e r . C 60 Q l a n d C 60 Q6 c o n t r o l s . All
transistors and all diodes do the same duty.
With sinewave drive it is more difficult to determine analytical formulas
for the mean and RMS currents in the transistors and diodes because the
duty-cycle required to achieve sinusoidal current varies in a nonlinear
manner throughout each half-cycle. Approximate estimates can be based
On the motor line currents. The peak transistor and diode currents are
equal to the peak line current. The mean transistor current is equal to

2-37
D e s ig n o f b r u s h l e s s p e r m a n en t -m a g n e t m o t o r s

or less than half the mean line current (since each transistor conducts
only on alternate half-cycles). The RMS transistor current is equal to or
less than 1 //2 of the RMS line current, or 1/2 the peak line current.
The mean and RMS diode currents are less than the transistor currents
under most condidons.

2.13 Maximum AC sinewave voltage available from DC supply


The torque production in the sinewave motor is based on the phasor
diagram (Chapter 6) which in turn is based on the fundamental voltage
and current. For a given DC supply voltage Vt, the RMS value of the
fundamental component of line-line voltage available to the motor is
^LLl(rms) , and the maximum value of this voltage depends on the control
strategy. For three-phase inverters the maximum value for various control
strategies is summarized in Table 2.9, using formulas developed in [2].
Another strategy that is sometimes used to increase the available AC
voltage while maintaining linear PWM control is to modulate the
switching with a third-harmonic component in the reference current
waveform in the current regulator. The potential of the neutral point
oscillates at third-harmonic frequency. With a three-wire connection, no
third-harmonic current flows in the lines.

Maximum available ^LLi(nro) V LLl(rms)

Six-step operation with 180* conduction in each 0.78


transistor 7T S

"Six-step1operation with 120 conduction in each 0.675


transistor ir fl
lim it of linear modulation in "sine/triangle" PWM y 0.612
2 i/2 S
T a b le 2.9 M a x im u m a v a il a b l e t h r e e - p h a s e r m s l i n e - l in e
FUNDAMENTAL AC VOLTAGE FROM A FIXED DC SUPPLY [2]

2-38
2. MOTOR AND CONTROLLER TYPES
In the single-phase full-bridge inverter (Fig. 5.16) the maximum RMS AC
output voltage is (2^2/n) Vs = 0.90 Vs in squarewave mode. At the limit
of linear modulation in "sine/triangle" PWM, it is V^/V2 = 0.71 Vt.

2.14 Provision of DC supply voltage


When the DC supply is rectified AC, the maximum value of Vs depends
on the type of filter used on the DC side. With a large filter capacitor on
the DC side and no inductors on either side of the rectifier, can in
principle approach the peak-peak. value of the AC line voltage. With a
single-phase AC line voltage of RMS value Va c this is expressed as =
V2 VAC = 1.41 For example, with V^c = 115V, = 162 V.
With a three-phase rectifier supplied from an AC voltage of V^L RMS
volts line-line, the maximum DC voltage with capacitive filtering is ^ l l A
For example, with = 460 V, Vt = 650 V.
With capacitive filtering at light load the AC line current is discontinuous
and highly distorted. Line inductors and increased load help to reduce
the AC line current distortion but they also reduce the DC voltage: with
constant DC current the voltage decreases theoretically to (3/ 2 /tt ) *1 l
= 1.35 ^LL, or 621V at 460 V AC.
The inrush current at switch-on, the AC line current harmonics, the
power factor, and conducted and radiated electromagnetic interference
(EMI) are all matters of rapidly increasing concern, and more and more
subject to regulation and legislation. Electromagnetic compatibility
regulations currently being developed and applied in Europe, the US
and elsewhere also cover the susceptibility of electronic controllers to EMI
generated in other equipment.

2.15 Controller architecture


2,15.1 Cost issues
The power section of a drive is very low in cost for small brushless motors
used in such applications such as disc drives and fans. The control
section can be of equal cost if not greater in some instances.
2-39
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

At higher power levels the cost of the higher power semiconductor


switches goes up much Easter than the extra ICs needed for more
sophisticated controls, (microcontrollers etc). At some level such as 1/2
hp the control secdon cost becomes relatively fixed and the increase in
inverter cost with power level is dominated by the cost of the
semiconductor switches, protection circuits, heat sinks, capacitors, and
current sensors. This cost structure is true for both sinewave and
squarewave drives, so that in higher-powered systems the complexity and
added cost of the sinewave drive is less significant. In low-powered
brushless systems the extra cost of the sinewave drive is a disadvantage.

AC R E C T IF IE R

POWER
IT C H I N G

CONTROLLER
OR M IC R O

EXTERNAI________ 1 SENSORS
COMMANDS

Fig. 2.20 Brushless drive block diagram

2.15.2 Squarewave drives


To supply DC variable voltages and currents and phase firing control for
the brushless permanent-magnet motor, the electronic controller requires
three basic blocks or sections. The first is a power supply which consists
of a power rectifier to convert the AC line voltage into a DC bus voltage.
2-40
2. M o t o r AND CONTROLLER TYPES
The second section is the inverter which includes the power switches
(usually two per phase) and their current sensors and protection
circuitry. The third section is the control, which decodes shaft angle
input data, controls DC voltage chopping (PWM), operates control loops
such as speed and position. Fig. 2.20 shows the relationships between
these blocks.
It was stated earlier that the power circuit of the electronic controller is
a switchmode circuit. The only means of controlling such circuits is to
control the timing of the gate signals that turn the power transistors on
and off. These low-level" timings must be controlled to meet a set of
"high-level" functional requirements. Working upwards, so to speak, from
the lowest level, the functional requirements are as follows:
1- Switch the power transistors on and off to regulate the
current to a predetermined value called the current reference
or set-point current 7sp.
2. Provide a means of protection against overcurrent in case
the motor stalls or has a short-circuit fault.
3. Synchronize the conduction periods of power transistors so
that the direction of current in each motor phase is
coordinated with the rotor position. This may include the
variation (phase advance) of the conduction periods to
overcome inductive effects particularly at high speed.
4. Determine the set-point current 7sp as a function of the
speed error signal A(i), which represents the difference
between desired speed and actual speed. Taking advantage
of the linearity of most brushless permanent-magnet motors,
this function may be no more than a simple
proportionality, i.e. 7sp A u. The constant of
proportionality is the gain. In servo systems the set-point
current has two terms: one proportional to the speed error
itself, and one proportional to the integral of the speed
error (PI control)- The function of the integral term is to
eliminate steady-state speed error, i.e. to force the
controller to make A to = 0 so that the actual speed is
exactly equal to the set-point speed. Without the integral
term, the only way to make the steady-state speed error
2-41
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

zero would be to use infinite "proportional gain", but this


is not practical for stability reasons. Sometimes even the
derivative of the speed error is also added. (See Chapter
14). The speed error signal may itself be derived from a
position-error signal in a position-control system.
The implementation of these functions for a squarewave drive is
represented in the block diagram, Fig. 2.21. The details of the individual
blocks in the diagram vary between manufacturers, but fundamentally
they all fulfil the same functions.

Vs Power electronics
MOTOR
Current
DC M mj} sensors Tacho
Driver
circuits

Shaft'
position
sensor
Current
regulators'^ I 1 j
Commutation
+ Logic

Speed Monitoring
Ref. err r - Motor tem p
speed -Supply voltage
l2 t
-Tacho
PI gain

Speed feedback

Fig. 2.21 Control system block diagram for squarewave brushless motor controller.

The commutation and current-regulation functions are both performed


by appropriate timing of the switching signals supplied to the power
transistors as described earlier.
2-42
2. M o t o r AND CONTROLLER TYPES
In a current-regulated drive, there is only one current reference signal
and this is used for all the six possible pairings of transistors (one upper
and one lower). The current reference signal is derived from the speed
error signal, which represents the difference between the desired speed
and the actual speed. The PI gain (proportional + integral gain) is
adjusted to optimize the control performance, (Chapter 14). The speed
reference signal is commonly an analog signal 0-10V representing zero
to full speed, and many drives are controlled from PLCs (programmable
logic controllers) with analog outputs. Many newer drives use digital
speed reference signals.
It is normal in manufactured drive products to include internal and
external monitoring and protective functions. Some examples are shown
in Fig. 2.21motor temperature, supply voltage, speed, and F t
(representing the accumulated heating effect of the windings).
Some excellent descriptions of particular systems are given in
manufacturers application notes, e.g. [3 ]. T h ere are several examples
of integrated circuits that comprise the entire control system of Fig. 2.21
on one chip, including some that operate without shaft position signals.
These "sensorless" circuits detect the rotor position from the back-EMF
waveform or from the variation of motor inductance with rotor position.
A particular implementation of the current-regulation strategy with fixed-
frequency voltage-PWM is shown in Fig. 2.22. The speed error signal,
which is an analog voltage, is compared with a sawtooth waveform whose
frequency is the switching frequency. The zero-crossings determine the
gating signals for the chopping transistor. If the speed error increases,
the duty-cycle increases. This increases the average voltage applied to the
motor increasing the current and the torque. Consequently the motor
accelerates and the speed error decreases. The proportional and integral
gains could be implemented by the feedback circuits of the op-amp
speed error amplifier. Alternatively they could be im plemented digitally
and then the digital speed error signal passed through an A-D converter
before being compared with the sawtooth waveform.
Rotation in the reverse direction is a simple matter of reversing the
torque at every rotor position, and this can be achieved in the
squarewave drive by "reverse commutation", i.e. by switching the
complementary transistor instead of the original transistor in Fig. 2.17.

2-43
D e s ig n o f b r u sh l e s s pe r m a n en t -m a g n e t m o t o r s

Speed error signal

^ State of chopping transistor

Fig. 2.22 Modulation of PWM duty-cyclc by error-amplificr signal.

Thus Q4 would conduct during the Q1 conduction interval; Q5 during


the Q2 interval, and Q 6 during the Q3 interval.
Braking or regenerating torque can be produced by the same method,
since the torque is proportional to current and the direction of the
current is controlled according to the rotor position, not the speed.
However, this presupposes that the controller remains in control of the
currents. At high speed (above the no-load speed), the back-EMF may
exceed the DC voltage to such an extent that the diodes act as a three-
phase bridge rectifier, even if the transistors are never switched on. The
DC filter capacitor may then be overcharged unless a braking resistor is
connected, as in Fig. 2.21.
A motor that can produce positive (motoring) torque and negative
(braking) torque in either direction is said to operate in four quadrants.
The ability to operate controllably in four quadrants, and even at zero
speed, is one of the hallmarks of servo motors.
If the reverse power flow is dissipated in a braking resistor, the braking
is called dynamic braking. If it is returned to the supply, the braking is
called regenerative braking. Dynamic braking is generally used for rapid
2-44
2. MOTOR AND CONTROLLER TYPES
braking, whereas regenerative braking is intended for overall energy
efficiency.
Another method of dynamic braking is to short-circuit the motor
windings while the motor is rotating at speed, or to connect them in
series with a wye- or delta-connected three-phase resistor. This can be
implemented with a relay that first disconnects the motor from the
controller. The motor generates power that is dissipated partly in the
external resistor and pardy in the motor windings. The torque depends
on the value of the external resistance. Quite high values of braking
torque and rapid deceleration are possible by this means. One
application is in the rapid deceleration of a washing machine tub if the
door is opened while the machine is running.
From Fig. 2.21 and 2.22 and the discussion preceding, it can be seen that
the shaft position signal is used by the squarewave drive for triggering
events, in particular the commutation of the current between lines. The
squarewave drive does not need a measure of the shaft position, either in
analog or digital form. Accordingly the integration of the shaft position
signals with the current-regulation or PWM signals can be implemented
purely in combinatorial logic, i.e. essentially in flip-flops. This is an
important distinction compared with the sinewave drive. In principle it
means that the simplest squarewave drives can operate with Hall-effect
sensors or optical interrupters that produce one pulse per commutation,
i.e. six pulses for every 360 electrical degrees (1 cycle) of rotation, on
three channels with two pulses per cycle per channel. The lowest-cost
brushless motors use this system.
Of course the squarewave motor is often used with an incremental
encoder, which produces typically 1000 pulses per revolution. The
commutation signals can be obtained from this pulse train also by
combinatorial logic. The reason for using a 1000-line encoder is to give
a better speed signal at lower speeds, if the speed is derived by a
frequency-to-voltage conversion circuit whose input is the encoder pulse
train. At zero speed or extremely low speeds, this method of speed
control does not work and the drive effectively needs to become a
position controller, and then it may be preferable to use a resolver
instead of an incremental encoder. Very low-speed operation normally
requires very smooth control and smooth torque production and
sinewave drives are usually preferred in such cases, although they may be
more expensive.
2-45
D e s ig n o f b r u sh l e s s p e r m a n e n t - m a g n e t m otors

2.15.3 Sinewave drives


The basic theoretical advantage of the sinewave drive over the
squarewave drive is very low torque ripple (less than 1%). This is widely
assumed to be an absolute fact but it is true only if certain conditions
are maintained: in particular, the waveforms of back-EMF and phase
current must both be sinusoidal.
Ideally the sinewave drive must control each phase current separately,
because the torque per ampere of each phase in a sinewave motor is a
sinusoidal function of rotor position. This requires separate PWM control
for the individual line currents. As the speed increases the motor
inductances have an increasing efFect and phase control of the current
becomes necessary to maintain the torque, (See Chapter 6).
Sinewave drives operate on a principle called vector control Although
there may be different definitions of vector control or field-oriented
control, all sinewave drives are fundamentally based on the two-axis
theory and the phasor diagram, Figs. 6.15-18. Vector control is
appropriate for sinewave motors in which torque is produced by the
interaction of the fundamental flux and the fundamental ampereconductor
distribution. Vector control generally operates with the d- and ^axis
current phasor components /d and /q, although these may be defined in
a variety of reference frames, either fixed to the rotor or fixed to the
stator. The determination of IA and / from the instantaneous line
currents requires a referenceframe transformation, which is a mathematical
matrix operation employing products and sums of current signals with
trigonometric functions of rotor position 0. In some cases the reference
frame transformation is implemented in a special IC called a vector rotator.
A reverse reference-frame transformation may be necessary to produce
the instantaneous sinusoidal set-point current signals for the current
regulator. Fully digital vector controllers may or may not use explicit
reference-frame transformations, but they are inevitably more complex
than squarewave controllers.
Vector control is described in detail in [4], which is recommended for
further study. This reference is particularly useful because it deals with
the vector control of the permanent-magnet synchronous motor in the
context of vector control of AC machines generally, so that the close
relationship with induction motor controllers can be clearly understood.
(See also [5-8]).
246
2. MOTOR AND CONTROLLER TYPES
If a sinewave motor is driven by a squarewave three-phase drive, the
theoretical torque ripple at low speed is at least 13%, even without taking
cogging torque into account. In principle the torque ripple can be
compensated by current-profiling and/or a high-gain velocity feedback
loop [9]. As speed is increased, it becomes more difficult to control
torque ripple because of limited controller bandwidth.
2.15.4 Unipolar drive
For minimum cost in low-powered systems it is possible to connect a
neutral connector or center tap to the wye connection and use only one
transistor per phase, as shown in Fig. 10.12. This circuit limits the phases
to only one direction of current, and the commutation frequency is one
half of that of a bipolar or full wave drive. The silicon cost is low but the
utilization of copper and iron is lower than with a bipolar drive.

2.16 Commutation test circuit


A very useful circuit for measuring the back EMF of a brushless motor
has been included for reference in Fig. 2.23. The purpose of the circuit
is shown by the example of the back EMF voltage generated from one
phase or pair of phases in series. On the same curve by the use of the
circuit the back EMF from all the phases is commutated using whatever
sensors are installed on the motor. This enables the designer to
determine exactly what the back EMF ripple wave shape looks like, and
it also indicates the accuracy of the commutation sensing scheme
whether it is included on the motor or whether it is remote sensing.

References
[1] Sebastian T, Slcmon GR and Rahman MA [1986] Design considerations for variable
tpnnd permanerd-magnet motorx, International Conference on Electrical Machines, Pt.
3, 1099-1102
[2] Mohan N, Undeland TM and Robbinj WP [1969] Pawn' electronics : converter*,
application* and design, John Wiley & Sons, NY ISBN 0-471-50537-4
[3] Motorola Application Notei MC330H [1990], MC33034 [1989]
[4] Vai P [1990] Vector control of AC. machines. Clarendon Press, Oxford ISBN 0-19-
859370-8

2-47
D e s ig n o f b r u sh l e s s p e r m a n en t -m a g n e t m o t o r s

Fig, 2.23 Commutated back-EMF current circuit

2-48
2. MOTOR AND CONTROLLER TYPES
Dote Y and Kinoshita S [1990] iJmiUrit servomotor*: fundamental* ami application*
Clarendon Press, Oxford
Leonhard W [1985] Control of electrical drive*, Springer-Verlag, Berlin
Vaj P [1992] Electrical machine* and drive*: a space-vector theory approach. Clarendon
Press, Oxford ISBN 0-19-859378-3
Boldea I and Nasar S [1992] Vector control of AC drive*, CRC Press, ISBN 0-8493-
4408-5
Jouve D and Bui D [1993] Torque ripple compensation inDSP-bavd brushless servo drive.
Intelligent Motion, PCIM Proceedings, Niimberg 28-37
3. BASIC DESIGN CHOICES
3.1 Introduction
Before a brushless motor design can begin, several important decisions
must be made. The reasons for this should be obvious from previous
discussion regarding the features of different types of brushless motors
and the availability of different magnetic materials. The method of
commutation is also an important issue which should be considered in
making the basic decisions about the design. The choice of an
interior rotor, exterior rotor, or an axial-gap motor must be made
first, along with a rough idea of the correct magnet grade. Then the
number of phases, the number of poles, the number of stator slots,
and the winding configuration must be selected. The rotor and
permanent magnet configuration is designed, then the stator and
winding are determined. A step-by-step procedure is provided as a
rough guide for a typical brushless DC motor design.
1. Determine application requirements (see Table 3.1)
2. Interior-rotor, exterior-rotor, or axial-gap configuration?
3. Select magnet grade
4. Select number of poles
5. Select number of stator slots and phases
6. Perform rough sizing estimate
7. Select air gap length and determine magnetic loading
8. Design rotor and determine flux/pole
9. Lay out stator lamination dimensions
10. Solve for numbers of conductors and turns/coil
11. Calculate wire size, resistance and inductance/phase
12. Calculate performance
13. Check temperature rise, current density, flux densities,
demagnetization of magnet
14. Modify design and reiterate until objectives are met
In this chapter the first five items are discussed, in terms of the basic
principles of brushless motor theory and practice. Later sections cover
rotor and stator designs, windings, and other practical aspects.
Aside from the motor design there are several other environmental and
performance requirements that must be taken into account. A checklist
of these is given in Table 3.1 for reference.
3-1
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

1 Appropriate national, US, EC, international & industry standards


2 Continuous power or torque requirement
3 Peak power or torque requirement
4 Maximum speed
5 Forward/reverse operation
6 Motoring/braking operation
7 Supply voltage
8 Supply frequency AC, or DC
9 Type of control required : torque, velocity, position
10 Precision and bandwidth required in closed-loop control
11 Programmability : motion profiles, slart/stop ramps etc.
12 Soft-starting requirements (inrush limitation)
13 Interface with PLC's, RS232, IEEE488, remote controllers etc. Front-panel
control requirements
14 Dynamic requirements: torque/inertia ratio, accel/dccel capability
15 Gearbox or direct drive
lfi Inlet Sc outlet temperature of available coolant; air/oil/water flow rate
17 Environmental factors : dust, hazardous chemicals, explosive gases.
Compatibility with insulation, magnets & other motor materials.
18 Maximum level of acoustic noise
19 Compliance with regulations on EMC and harmonics
20 Warranty requirements
21 Maintenance; spares
22 Operator's manual and repair manual
23 Vibration withstand levels
24 Fault protection: type of protection required (overcurrent, overvoltage,
undervoltage, over temperature, winding faults, vibration sensors)

T a b le 3.1 C h e c k l is t o f a p p l ic a t io n r e q u ir e m e n t s

3-2
3. B a sic d e s ig n c h o ic e s

3.2 Interior rotor, exterior rotor, or axial gap configuration?


If a fairly low speed, constant speed or slightly variable speed brushless
motor design is required, it would be wise to consider an axial rotor
design. This is particularly true if zero cogging and smooth operation is
required. Applications with this requirement include record turntables,
CD players, floppy discs and VCRs. In these applications the output
power is low and the speed is often under 1,000 rev/min.
If a high-torque, low-speed machine is required, then an interior-rotor
design would be appropriate using rare earth magnets and a high pole
count. Such motors can be made with a large hole through the center of
the rotor, which provides valuable space for other parts of mechanisms,
cabling, or cooling media.
If a higher speed is required which is constant or varies only slightly, an
exterior-rotor motor should be considered. The features of these designs
have to do with ease of manufacturing and low cost. The relatively high
rotor inertia is desirable for such applications as fans and blowers.
Manufacturing processes for brush-type DC motors are readily adaptable
to manufacturing exterior-rotor brushless motors. DC armature winding
machines can easily be adapted to wind the stators. The assembly and
adhesive techniques required for assembling the rotor magnets are well
suited to the production of exterior rotors for brushless motors.
Manufacturing processes for AC induction motors are more readily
adaptable to manufacturing interior-rotor brushless motors. If the
requirement is a servo with high performance for factory automation,
machine tools or aerospace there is little choice but to select the interior-
rotor design using either high-performance rare-earth permanent
magnets or an embedded ferrite rotor design. Either of these would tend
to require a completely new stator lamination to carry the high magnet
flux.

3.3 Number of phases


Brushless DC motors are often assumed to have three phases but this is
not always the case. One of the first brushless DC motors to appear on
the market in the early 1960s, was a 4-phase motor made by Siemens. It
required 8 transistors, 2 for each phase. This was not a serious problem
3-3
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a c n e t m o t o r s

because it was a veiy small, low-powered motor. 4-phase brushless motors


are not much used today. However, there are many 2-phase brushless
motors in existence. Fig. 2.6 shows a small brushless fan which is
manufactured in 3-phase and 2-phase versions. The 2-phase version is
popular because only one Hall switch is required for commutation and
only 2 power transistors are needed. This makes them cost-effective for
light-duty fan applications. The problem with 2-phase motors, sometimes
called single phase bijUar-wound motors, is that the torque passes through
zero every 180 electrical degrees. If by chance the rotor comes to rest at
a position of zero torque, the motor would fail to start the next time it
was turned on. Also, the direction of rotation is difficult to predict if the
magnets and stator laminations are symmetrical. To overcome this
problem various designs have emerged which typically use space-
harmonic fluxes that may vary under saturation of the iron, producing
enough of a shift in the zero-torque position to get the motor started [ 1].
For sinewave motors the minimum number of phases needed to produce
a pure forward-rotating ampere-conductor distribution with no reverse-
rotating component is two, and some servo motors are manufactured
with two phases. The phase windings are sine-distributed (see Chapter 6)
and their axes are orthogonal. They are fed with currents in phase
quadrature, and the rotation of the ampere-conductor distribution is
derived from the relationship
/ sin p9 cos <i>s/ + j sin (p9 - ) cos (v>st - )
2 2 2 2 (3J)

= / ,' 2 ? s in ( p O ~ c o s

where *is the peak phase current and Nt is the effective number of sine-
distributed turns per phase (Chapter 6). This can be compared with the
equivalent relationship for three-phase motors, equation (6.21). For full
control flexibility and maximum utilization of the available DC voltage,
each phase needs to be fed from a full-bridge inverter circuit (Fig. 5.16),
requiring a total of eight transistors and eight power diodes. This may
seem excessive compared to the six transistors needed in a three-phase
drive, but full-bridge single-phase inverters can be neatly packaged, and
the total Silicon area need be no more than in the equivalent three-
phase drive. With separate full-bridge inverters the two-phase motor also
requires four connecting leads, compared with only three for a three-
phase motor. The number of stator slots in the two-phase motor does not
3-4
3. B a sic d e sig n c h o ic e s

fit naturally into the standards used with three-phase motors, and special
laminations may very well be required.
Three-phase motors are by far the most common choice for all but the
lowest power levels. In common with AC motors generally, they have
extremely good utilization of copper, iron, magnet, insulating materials,
and Silicon, in terms of the quantity of these materials required for a
given output power. Although the utilization can theoretically be argued
to be higher in motors of higher phase number, the gains would be
offset by the increased number of leads and transistors, which increases
cost and may severely compromise reliability. The only practical
application might be for a brushless DC torque motor for direct-drive
applications where backlash from gear reducers is unacceptable.
Three-phase motors have the flexibility afforded by wye- or delta-
connected windings, or even unipolar windings. They can operate with
only three connecting leads with no loss of control flexibility. They have
excellent starting characteristics, with smooth rotation in either direction,
and low torque ripple. They can work with a very wide range of magnet
configurations and an enormous range of winding configurations, and
can take advantage of the coil-winding technology that has been
developed for both AC induction motors and DC brush-type motors.
They can operate with either squarewave drive or sinewave drive, and are
Well adapted to the development of "sensorless" controllers that require
no physical shaft position sensor.
No. of phajcj Conductor utilization No. of power Torque ripple %
% twitches
1 50 2 100
2 SO 4 or 8 30
3 67 fi or 3 IS
4 75 8 10
6 83 12 7
12 92 24 3

Table 3.2 C o m pa r iso n o f brushless m o t o r s w it h d iffer en t ph a se


nu m bers

3-5
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

Table 3.2 shows a comparison of the conductor utilization, the number


of commutation sensors, the number of power switches, and the resulting
torque-ripple percentages. When a brushless DC motor operates at very
low speeds (under 100 rev/min) the torque waveform has the same
variable content as that of the back EMF waveform, as in the overlap
diagram of Fig. 2.23. The torque ripple percentages in Table 3.2 are
based on sinusoidal back-EMF at low speeds. The conductor utilization
is equal to the actual number of conductors carrying current at any one
time in squarewave drive, as a percentage of the total number of
conductors available.

3.4 Number of stator slots and poles


The choice of the number of poles depends upon many factors, some of
which are as follows:
1. Magnet material and grade
2. Interior-rotor vs. exterior-rotor vs. axial-gap rotor
3. Mechanical assembly of the rotor and magnets
4. Speed of rotation
5. Inertia requirements
Before any detailed analysis of the number of poles is undertaken, a basic
nde of thumb should be considered. The number of poles should be
inversely proportional to the maximum speed of rotation. The reason, of
course, is to limit the commutation frequency to avoid excessive switching
losses in the transistors and iron losses in the stator. For very high speeds,
two- and four-pole motors are preferred. If smooth torque is required at
low speed, such as in a DC torque motor, a larger number of poles
should be selected.
Every time the number of poles is doubled the required thickness of the
rotor yoke or back-iron inside the magnets is reduced by one half, as is
the thickness of the stator yoke. Therefore, for a given magnetic and
electric loading and a given rotor diameter, the overall diameter can be
reduced by increasing the number of poles.
As the number of poles increases, the stator ampere-conductors per pole
decrease in inverse proportion, so that the per-unit inductance and
synchronous reactance decrease in motors of higher pole-number.
3-6
3. BASIC DESIGN CHOICES

8 POLE

SKEWED RING
MAGNET

Fig 9.1 Interior-rotor configurations, (a) 2-pole flux-concentrating embedded-magnet


design. This impractical unbalanced arrangement shows the difficulty of making
a flux-concentrating design with only 2 poles, (b) 6-pole "spoke" design with
radiused pole pieces, (c) Surface-magnet rotor with skewed magnets to reduce
cogging, (dj 4-pole surface-magnet rotor with magnets held on rotor surface by
adhesive, (e) 12-pole "spoke" design with retaining ring, (f) Ring magnet with
bonded or moulded magnet with multiple poles magnetized after assembly.
3-7
D e s ig n o f b r u sh l e s s pe r m a n en t -m a g n e t m o t o r s

Two-pole motors not only have the largest diameter, but they also have
the greatest susceptibility to magnetic unbalance which can lead to shaft
flux and induced currents in the bearings.
Another important point regarding the number of poles has to do with
cost. For example, if a bonded ring magnet is used which is molded
from one of the rare-earth materials or a ferrite material, it is easy to
magnetize any number of poles desired on the outside diameter or inside
diameter of the ring. In fact, it does not cost any more to magnetize two
poles than it does a hundred poles once the magnetizing fixture is paid
for. On the other hand, if the motor uses arcs or blocks of samarium
cobalt, the greater the number of poles the greater is the cost in magnets
and labor for fabrication.
It is not surprising then to find that most brushless motors have four, six
or eight poles, with four poles being the most popular choice.
If a rotor is to be designed using embedded slab magnets, there are
several possible configurations. It is easier to configure a design with low
flux leakage from the poles if a high pole number is used. Examples are
shown in Figs. 2.2 e, 3.1a, 3.16 and 3.1 e. It is clear from these examples
that a certain amount of flux can leak from pole to pole inside the rotor,
rather than crossing the airgap and interacting with the stator current to
produce torque. If the magnets are thick enough a six-pole design can
be used with embedded flat magnets, but in most instances an 8-pole
design is required to minimize the flux leakage on the inside o f the
rotor.
As shown in Fig. 3 . 1 skewing of the magnet poles may be used to
reduce cogging torque. This requires specially tooled magnet arcs. If a
ring magnet is used as shown in Fig. 3.1/ the poles can be skewed
magnetically in the magnetizing fixture. The other possibility to reduce
cogging is to radius or chamfer the edges of the magnet poles as shown
in Figs. 3.16, d, and e. This so-called pole-shaping at the edges of the
poles can be more cost-effective because specially tooled skewed pole arcs
are difficult to manufacture. The actual pole arc angle itself will be
covered in the next section.

3-8
3. BASIC DESIGN CHOICES
Slots 8 IS IB 20 24 28 32 36 40 44 48
Poles 2 2 2 2 2 2 2 2 2 2 2
6 10 4 6 4 6 4 6 4 fi 4
fi 14 6 10 fi 10 fi 10 fi
10 10 18 8 14 10 14 8
12 18 22 10 22 12 18 10
14 20 12 26 14 12
14 30 26 14
20 28 18
22 30 20
24 34 SO
26 34
28 36
38
40
42

T able 3.3 S l o t / P o l e c o m b in a t io n s f o r 2-ph a se m o t o r s

Until recently, it appears that certain pole and slot numbers are more
popular than others. However, there are many combinations of slot- and
pole-numbers that can be used effectively. Tables 3.33.7 list all the
possible pole-numbers which will operate with stator laminations having
slot-numbers from 3 to 48, for 2, 3, 4, 5, and 6 phases.
Many of the combinations listed are not immediately obvious from a
casual observation, and the tables have been generated with the aid of a
special computer program. It is quite possible that there are useful
applications for some of these slot/pole/phase combinations that have
not been previously used.

3-9
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

Slot) 3 6 9 12 15 18 21 24 27 SO 33 Sfi 39 42 45 48
Polci 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2
4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4
6 a 10 6 8 8 6 8 8 6 8 8 6 8
8 10 8 14 10 8 10 10 8 10 in 8 10
12 18 12 16 16 10 20 14 10 14 14 10 14
14 20 12 22 20 12 16 16 12 16
16 18 26 22 14 26 26 14 20
20 26 16 28 28 16 32
22 28 22 32 32 20 34
24 24 34 34 28 38
26 30 40
28 32
30 34
32 38
40

T able 3.4 Sl o t / P o l e c o m b in a tio n s fo r 3-ph a se m o t o r s


Perhaps one of the most common brushless 3-phase motors has 4 poles
and a 12-slot laminadon with a lap winding having coils of 3-slot pitch.
This design tends to have a high cogging torque and large end-turns in
the stator which contribute to /iZ losses. For interior- rotor designs this
winding is difficult to wind without the use of expensive machinery.
Table 3.8 shows some of the most popular slot and pole combinations for
3-phase motors which have been categorized in terms of their slots/pole
ratios. This list or grouping provides a clear summary of the choices
possible for slots and poles for brushless motors in terms of how well they
will resist cogging because of pole to slot alignments. This works out
equally well for interior-rotor, exterior-rotor, or axial-gap machines.
Perhaps the most important point about these possibilities is that some
of them are fractional-slot designs and some are integral-slot designs.
This is significant in relation to the inherent cogging torque. (See also
section 4.8).

3-10
3. B a sic d e s ic n c h o ic e s

Slotl 8 16 24 32 441
Polei 2 2 2 2 2 " l
2
4 4 6 4 6 4
6 10 6 10 6
10 18 8 14 10
12 10 26 12
14 12 SO 14
14 34 18
20 20
22 30
24 34
26 36
28 38
40

T a ble 3.5 Sl o t / P o l e c o m b in a tio n s fo r 4-ph ase m otors

Slots 5 10 15 20 25 SO 35 40 45 |
Pole. 2 2 2 2 2 2 2 2 2 i
4 4 4 4 4 4 4 4 4
6 6 6 6 6 6 6 6
8 12 8 8 8 8 8 8
12 10 12 12 12 12
14 16 IB 14 14 14
Ifi IB 22 22 16 16
20 24 24 24 IB
22 26 26 26 28
2B 28 32
32 34
34 36

T a b l e 3.6 Sl o t / P o l e c o m b in a t io n s f o r 5-p h a s e m otors

3 -n
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

Slots 12 24 3fi
Poles 2 2 2
10 4 fi
10 10
20 14
22
26
30

T able 3.7 Sl o t / P o l e c o m b in a t io n s f o r 6-p h a s e m otors

For example, if the slots/pole is even, then every edge of every pole lines
up with every slot, causing cogging. If a fracdonal-slot combination is
used, fewer pole-edges line up with the slots. The actual pole arc can
make this situation worse or better. A fractional-slot configuration
minimizes the need for skewing of either the poles or the lamination
stack to reduce cogging. In fact, it is recommended that a integral-slot
winding never be used. Note that this precludes one of the most famous
brushless motorsthe one with 4 poles and 12 slots, as well as all of its
derivatives from the 3 slots/pole series.
A final point to be made about the slot and pole relationship concerns
the winding pitch. Since the coils can be wound only over an integral
number of slots, the winding pitch is determined by dividing the number
of slots by the number of poles and rounding off to the next lower
number, or in the case of the 0.75 slot/pole series, the next larger whole
number. The winding pitch or span is summarised in Table 3.9 for all
the slots/pole ratios in Table 3.8.
It should be obvious that the end turns are shortest when the pitch is
one or two slot-pitches. Anything above 2 requires a considerable
overlapping of end turns from one coil around the preceding coil end
turn. This requires that coils be either hand-inserted or machine wound
with expensive AC induction-motor winding equipment. The single-slot
and 2^slot pitch windings can be automatically wound on needle winders
used for winding stepping motors and series motor stators, and are
economical to make.
3-12
3. B a sic d e s ig n c h o ic e s

0.75 1.125 1.5


Slots Poles Slots Poles Slots Poles
3 4 9 8 3 2
6 8 18 lfi 6 4
9 12 36 32 9 6
12 16 12 8
15 20 15 10
18 24 18 12
21 28 21 14
24 32 24 16
2.25 3.75
Slots Poles Slots Poles Slots Poles
9 4 6 2 15 4
IB 8 12 4 30 8
27 12 18 6 45 12
24 8
30 10
36 12
4.5 5.25 6
Slots Poles Slots Poles Slots Poles
9 2 21 4 12 2
18 4 42 8 24 4
27 6 36 6
36 8 48 8

T able 3.8 S l o t / P o l e ra tio s f o r 3-ph ase bru sh less m o t o r s f o r


m in im u m c o g g in c t o r q u e

3-13
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

Slots/pole Coil span in slot-


pitches
0.75 1
1.125 1
1.5 1
2.25 2
3 3
S.75 3
4.5 4
5.25 5
6 6

T able 3.9 C o il span o r p it c h f o r d iffer en t sl o t / p o l e ra tio s

If the brushless motor is being evolved from an AC induction motor


family, then the winding pitch is not an issue because high-speed stator
winders would be available. These motors will have winding pitches
similar to those of induction motors in order to utilize the equipment,
and much of the preceding discussion is not applicable.

3.5 Caution with the basic laws of electromagnetism


In this section we take a breather from the design trail and pause to
reflect on the laws of electromagnetism as applied to the design of
electric machines. The discussion is not actually necessary for readers
wishing to keep going with the design story, but it is placed here (and
not in an appendix) because it is felt to be necessary to the rigour of the
design equations that follow, indeed of the entire book.
Several basic laws of electromagnetism are fundamental to the operation
of all electromagnetic motors and generators. The first is Faradays Lata
which states that the voltage induced in a circuit is equal to the rate of

3-14
3. B a sic d e s ic n c h o ic e s

change of magnetic flux-linkage.1 The concept of flux-linkage is


conveniently thought of as the product of the number of turns and the
flux going through them, although this is not exact because the flux is
a diffuse distribution that may not link all the turns equally. The change
of flux-linkage can be caused either by rotation of the magnet past the
coils of the circuit, or by a change in any current that is contributing to
the flux linked by the circuit. The voltage induced by rotation is called
a rotational voltage. This is the nature of the back-EMF or open-circuit
voltage. A voltage induced by a changing current, either in the circuit
itself or in another circuit nearby, is sometimes called a transformer EMF
but it is better to think of this type of voltage as being associated with self
inductance (Z, di/dt) or with mutual inductance (M di/dt).
The term "EMF" describes the circuit as an electrical source or generator.
With this convention the correct statement of Faradays Law is e -
-di|/dt The minus sign is sometimes elevated to the status of a law in
its own rightLenzs Law. This states that the current (that would flow if
driven by e around the circuit) will tend to establish a flux in the
opposite direction to that of the flux that is causing the EMF in the first
place. The source or generator convendon is illustrated in Fig. 3.2along
with the corresponding direction of positive current flow for generator
action.
In electric motors the voltage at the terminals is applied from a separate
source or power supply (the electronic drive), which forces the current
to flow into the windings in the positive direction for motoring as shown
in Fig. 3.26. The motor winding is regarded as a load or sink, and the
back-EMFir regarded as a voltage-drop. Accordingly the correct statement of
Faradays Law is e - d|f/d( without the minus sign. In motor calculations
the minus sign would be a nuisance likely to lead to errors, and in this
book the motor or sink convention is used.

'American readers will be interested to know that Joseph Henry, an Albany, N.Y.,
schoolmaster for whom the unit of inductance is named, discovered the law of
electromagnetic induction at about the same time as Faraday, if not before, but Faraday was
narrow!)' first to publish. Ironically, Faraday gave his name (or part of it) to the unit of
capacitance, which is of very liltlc significance in electrical machines! One of the reasons
for Faradays experimental success, after famous scientists had searched in vain for
electromagnetic induction for many years, was that as a chemist he took much trouble to
Obtain and use high-conductivity oxygen-free copper for winding his coils, which at the time
was not generally available. He also owned stock in a Welsh copper mine.
3-15
D e s ig n o f b r u sh l e s s per m a n en t -m a g n e t m o t o r s

i 1

Positive Positive
generator motor i
current current

Generated EMF Back-EMF

a. SOURCE convention b. SINK convention

Fig. 3.2 Source and sink convention!

Similar consideration applies to the voltage drop in an inductance L,


which is treated as = L di/dt without the minus sign. Both the "back
EMF' and the voltage drop in the winding inductance can be added to
the resistance voltage-drop Ri> giving an equation for the terminal voltage
v= Ri + L di/dt + e (plus any mutual-inductance terms: see Chapter 10)
in which all the terms are positive. With this convention e should really
be called the rotational voltage-drop but the term back-EMF is so well
established that this would only cause confusion. So in this book we use
the term back-EMF but we omit the minus sign, and treat it as a voltage
drop.
The second basic law of electromagnetism is the law offorce on a current-
carrying conductor, F= BIL, where B is the flux-density, 7 is the current,
and L is the conductor length. This law is the magnetic term from the
Lorentz force equation (with the electrostatic term omitted). In slotless
motors there is apparently no difficulty in applying this law; but in
motors with slotted iron stators the conductors are laid in slots and the
magnet flux goes through the Uethr-not through the conductors at all,
making it uncertain whether the BIL formula can be validly used.
The widespread acceptance of the BIL formula is even more surprising
when one questions where the force acts. Presumably it acts on the
copper conductor, yet the current is carried by the moving electrons. To
explain this it is necessary to invoke the Hall effect which says that the

3-16
3. B a sic d e sig n c h o ic e s

electrons drift to one side of the conductor. The force really appears on
the electrons but since they are, on average, displaced from their
electrically-neutral distribution, there is a force of attraction between
them and the positively-charged ions of the copper lattice, in a direction
across the conductor orthogonal to the direction of current flow. By this
means the force is transmitted as an electrostatic force to the conductor
itself. The Hall coefficient, which relates the electrostatic voltage across the
conductor to the product of the magnetic field and the current, is
extremely small in copper, which is probably why this effect is rarely
discussed or even noticed.
The Lorentz force law is unnecessary in deriving motor design equations
because the desired equation for the torque can usually be derived from
Faraday's Law with the law of conservation of energy, embodied in the
principle of virtual work and the concept of coenergysee Chapter 5. The
widespread use of the Lorentz law is undoubtedly due to its simplicity. In
many cases it gives the correct result, but in slotted motors this is
fortuitous and it is always wise to question its validity.
The force law, however derived and formulated, becomes a torque law
in rotating machines because the rotor is mounted in bearings and the
tangential force "appearing at the rotor surface" moves with the rotor.
The most fundamental laws or equations involved in the design and
performance-analysis of electrical machines are summarized in Table
3.10, along with the basic laws of electromagnetism that go with them.

EMF equation Faraday's law


Torque or force equation Conservation of energy combined with
Faraday's law; principle of virtual work
and coenergy. Lorentz law
Electric circuit equation Kirchoffs law
Calculation of magnetic flux, flux-density, Gauss law, Amperes law and the
flux-linkage and inductancc constitutive relationships between B and
H for steel, air, and magnets
T able 3.10 L aws o f electr o m a g n etism in electr ic a l m a c h in e d esig n

2 It was fint pointed out to the author by D.W. Jones,

3-17
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

3.6 Simplified motor design


After the basic choices of configuration, pole-number, etc. have been
made, it is advisable to perform a detailed design calculation using
various grades of magnet, on which a cost comparison can be based.
The main goal in designing the PM rotor is to provide the maximum
number of flux lines per pole $ in such a configuration as to cost as
little as possible (in terms of magnet cost and assembly labor), with the
minimum amount of flux leakage. Before calculating the flux, however,
it is advisable to understand the EMF and torque constants which link
the electrical, the magnetic, and the mechanical aspects of the design.
3.6.1 Simple design formulas for the EMF and torque constants
As the PM rotor rotates, the flux-linkage of a phase winding varies from
a maximum positive value $ to a maximum negative value - 0 . The
transition takes place over n electrical radians. An electrical radian is
equal to a mechanical (actual) radian divided by the number of pole-
pairs in one revolution, and if the motor is rotating at <i)m actual radians
per second, the transition tunifrom to 4> is n/<i>mp seconds. If the
complete machine (three phases) has Z conductors, then there are 1/3
x Z/2 turns/'phase, since one turn comprises two conductors.
Consequently, by Faradays law, with two phases in series at any time
(squarewave motor), the average back-EMF over half an electrical cycle (it
electrical radians or 180 electrical degrees) is
E = 2* ~ * 1 . 2 Z& p = . (3 2)
ir /a mp 2*3 s 3 air m
in [V], where a is the number of parallel paths into which the phase
winding is divided. The average back-EMF is therefore proportional to
the speed, and the constant of proportionality AE is the back-EMF constant.
In equation 3.2 it is expressed in volt-sec/rad, but it is often measured
and quoted in Volts/krpm, i.e. volts per 1000 rev/min.
Care is required in using equation (3.2), for the following reasons:
The effective number of conductors Z must be determined
for each situation. Because of the winding distribution, not

3-18
3. B a sic d e sig n c h o ic e s

all of the turns link the maximum flux 4> at the same time,
and their contributions to the total winding EMF are
generally not in phase with each other and should be
summed vectorially.
The magnets do not generally produce a perfect
squarewave of flux in the airgap. As a result, the average
EMF calculated by equation (3.2) is not equal to the peak
EMF, even if all the turns are concentrated together.
The back-EMF constant can be modified to take these practical factors
into account by writing
Jc = l? ? p c. (3.3)
3 an
The value of the coefficient C depends on the pole arc and the winding
distribution and connection. The effects of these are discussed at length
in later chapters, particularly Chapter 7. For delta connection, the 2/3
factor is replaced by 1/3.
If $ is measured in SI units (Wb or Webers), and in mechanical
rad/sec, then is in volts. (Z, p, and a are numeric with no dimensions).
If <E> is measured in lines, then equation (3.3) should be multiplied by
10 8 to get E in volts.
As shown in chapters 1 and 5, in squarewave motors the torque is
proportional to the DC current / and the constant of proportionality or
torque constant fcj- is equal to fcE if both are expressed in SI units (Nm/A
and V-s/rad). Other units are frequendy quoted, however, and for
reference the relationships between some common units are summarized
in Table 3.11.

8 Note that the RMS value is different again. Unlike the RMS value of the current,
the RMS value of the back-EMF is of no interest in brushleu motor design. However, care
should be taken while measuring it with electronic instruments (particularly multimeters),
because these arc often set up to measure and indicate RMS values and not average or peak
values. The way to be certain of back-EMF values is to measure them on an oscilloscope,
preferably a digital processing oscilloscope that can calculate the mean value while
displaying the whole waveform. The waveform is of interest for many other reasons, not
least for analyzing torque ripple.
3-19
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

1 Nm/A - 141.fi oi-in/A - 8.85 in-lb/A


1 V-s/rad - 104.7 V/krpm

T a b le 3.11 C o n v e rs io n e q u a tio n s f o r u n its o f kr a n d


It is perhaps worth noting that the flux $ in the expressions for the EMF
and torque constants is the magnet flux acting alone on open-circuit, i.e.
with no stator current flowing. It does not include any contribution from
the stator currents. Although the stator currents do produce additional
flux ("armature reaction"), this is accounted for by the winding
inductances (self and mutual), and it does not contribute to the torque
unless the rotor is of the salient-pole type (Chapter 6) or there is very
heavy saturation of the stator.
A more detailed analysis of the definition and use of and AE is
provided in Chapter 7. The most important point here is to recognize
the importance of two variables in both equations. Z and $ are the only
adjustable variables once the basic design decisions have been made such
as the number of phases, poles, slots, coils/phase and winding
arrangement. $ is the total number of useable lines of magnet flux per
pole produced by the magnet acting alone (i.e. on open-circuit), and
essentially comes from the rotor design including the magnets. The
number of conductors Z is equal to the product 2 x (tums/coil) x
(coils/phase) x phases, which comes from the stator design. It is
convenient that the torque and voltage formulas have no dimensions in
them to worry about, otherwise the number of variables would be
inconveniently large and the design calculations could become too
reiterative.
3.6.2 Simple calculation of the flux
The simple classical method for calculating the flux $ is to start with
the B /H curve for the magnet, as supplied by the permanent magnet
vendor. A typical curve is shown in Fig. 4.5 and the constructions
needed here are shown in Fig. 4.4. If a "hard" magnet is used in the
form of an arc of uniform thickness and uniform magnetization, the
resulting flux lines across the air gap should be uniform except where
the stator slot openings are located. The first step is to determine the
permeance coefficient PC which characterizes the operation of the

3-20
3. B a sic d e s ig n c h o ic e s

magnet within the magnetic circuit formed by the motor geometry, using
equadon (4.12):

PC = * h* x A (3.4)
f lX G g AU

where g / is the effective airgap (taking Carters coefficient for slot-


openings into account), LM is the magnet length in the direction of
magnetization, AM is the magnet pole area, A is the pole area at the
airgap (nearly equal to in surface magnet m otors), and J lkc is a
leakage coefficient with a typical value of 0.80.9. For flux-concentrating
designs such as the "spoke" design, this equation should be carefully
worked out with regard to the number of magnets contributing to the
flux through each pole at the airgap. (See Chapters 4, 6 and 13).
If the arc-tangent of PC is computed on a hand calculator, the angle of
the load line from the negative *-axis can be determined, and the load
line should be plotted to find the intersection with the magnet
characteristic (Fig. 4.4). This is the open-circuit operating point of the
magnet. It is best to use magnet data curves which show several different
temperatures. For example, if the expected continuous operating
temperature of the magnet is 100C then the 100 C curve should be
used if it is available. Otherwise, if the intersection of the load line and
the magnet characteristic is determined at 25 C then the number of
stator conductors Z would need to be increased by multiplying it by the
product of the temperature rise above 25 C and the temperature
coefficient of the magnet: in particular, the temperature coefficient of Br
The y-axis value of the intersection of the load line with the magnet
characteristic is the open-circuit operating flux-density of the magnet, 5 M.
Note that it is lower than the remanent flux-density Br, which would be
obtained only with zero airgap, infinitely-permeable steel, and no leakage.
Depending upon the magnetic circuit configuration, the leakage ( /l^ )
can vary considerably. The factor JlkG is usually only an estimate:
accurate calculation would require tedious finite-element calculations and
this is in many cases not worth the effort. The best remedy for
maximum utilization of available flux from a magnet is experience and
common sense. If performance (kj) is a bit low when testing the

3-21
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

prototype, the addition of a proportionate number of turns per coil


should fix the problem. If Aj- is a bit high, take off turns.
To compute the flux/pole $ for insertion into the equations for kj. and
kE, the pole area of the magnet is multiplied by the open-circuit flux
density
* = <3-5)
taking into account all magnet poles that contribute flux to the airgap
pole (in "spoke" designs, there are two magnet poles per airgap pole).

[ N o te : The term energy product (measured in MGO or mega-Gauss-Oersteds)


is often used as a measure of the "strength" or grade of magnets,
especially when a new grade is announced. In modern "hard" magnets
with a straight demagnetization characteristic, in which and Br are
practically equal, the energy-product serves as a single number that
characterizes both the coercivity and the remanence. In former times,
before the introduction of "hard" magnets, Alnico magnets were
popular. Since the coercivity of Alnico (in Oersteds) is numerically only
1/10 to 1/20 of its remanence (in Gauss), it was necessary to design the
magnet to operate at a load line well above the point on the magnet B /H
curve where the energy-product was a maximum, particularly if the
demagnetizing MMF of the armature current was significant. This
approach to magnet design has long been viewed as yielding the most
cost effective design.
Hard Ferrite and Rare-earth magnets available today have such high
coercivity that demagnetization is not such a problem, and they can
operate with much lower permeance coefficients. This is what permits the
use of flux-concentrating configurations like the "spoke" design, which
would not be possible using low-coercivity Alnico magnets. In surface-
magnet motors it is good practice to design with a high permeance
coefficient, even with high-energy, high-coercivity magnetsnot only to
maximize the flux, but also because the difficulty of making magnets thin
enough to operate at maximum energy-product can result in higher
magnet cost.]

3-22
3. BASIC DESIGN CHOICES
3.6.3 Basic sizing rules
Magn& overhangIf Ferrite magnets are used as shown in Fig. 2.5, it is
normal practice to increase the magnet area and resulting flux by making
the magnet axial length longer than the stator lamination stack. The
effects of increased flux by using magnet overhang can be quite
significant on a shores tack motor. The amount of overhang should
never exceed the magnet thickness on each end. Surface-magnet rotors
with Rare-earth magnets normally do not use overhang because the
remanence is so high that the stator teeth cannot cany the extra flux,
their width already having been made as small as possible to maximize
winding space.
Qmsequentfole deignsOne of the main advantages of consequent-pole
designs (as shown in Fig. 2.2d) is that operation at the maximum energy-
product is practical because the thicker magnet is used to drive two air
gaps, which decreases the permeance coefficient.
*Spoke* designsIn the case of the "spoke" design, Fig. 2.2e, the magnet
thickness is again based on two airgaps in series, each with one half of
the soft iron pole area. This configuration is cost-effective for both
Ferrite and bonded NdFeB grades because the magnets are flat slabs.
They can be easily magnetized before assembly without concerns about
cleaning after machining. Given the accuracy of the soft iron pole pieces
between the magnets there is no need for machining or grinding on the
finished rotor assembly. Mechanical retention of embedded magnets in
the spoke configuration is more dependable than than any method
which relies solely on adhesive. Other benefits of the embedded-magnet
rotor (spoke and interior-rotor types) include low magnet cost, low
magnet tooling cost, and high airgap flux-density. Also the soft iron pole-
piece can be shaped to reduce cogging.

Airgap lengthMaximum resistance to demagnetization requires a small


airgap length. On very small motors, this can be 0.005" to 0.010" on a
side. On medium size motors 0.015" to 0.020" is sufficient. On large
brushless motors, 0.025 to 0.035 on a side is appropriate.
First estimate of magnet thicknessA rough-and-ready first estimate of
magnet thickness is about ten times the airgap length. (If there are two
airgaps per magnet, add them together). This method of first
3-23
D e s ig n o f b r u sh l e s s per m a n en t -m a g n e t m o t o r s

approximation applies to the use of high-coercivity magnets such as the


ferrites and the rare- earths.
Exploratory selection of magnet gradeDo not waste time with concerns about
energy-product. In general, select the magnet grades with the highest
remanence available at reasonable prices. Then perform an initial design
using a magnet thickness of ten times the air gap. Lay out the rotor to
scale no matter what the configuration, and take a copy of the normal
B fH curve at operating temperature.
Effective dimensions of arc magnetsIf an arc is used as a magnet shape, use
the minimum arc length to calculate the magnet pole area This is
always the inside arc if the configuration is surface-magnet with radial
magnetization. Contrary to popular belief, it is incorrect to use the mean
radius of the magnet arc to calculate the pole area.
Care (yoke or back-won) dimensionsThe magnet core, yoke or back iron as
well as soft iron poles should be sized to keep flux densities as low as
possible. Usually in an interior-rotor design the issue is not as serious as
in exterior-rotor designs. In exterior-rotor designs the steel return path
or "rotor cup" usually overhangs the magnet to provide extra cross-
sectional area for the flux return path, and this permits the use of a
thinner gauge of steel, Fig. 2.5. The flange of rotor cup also contributes
area to the flux return path.
Use of carbon steelflux return path-If carbon steel is used for a rotor cup
or any other form of magnet return path, it is advisable to keep the flux
density under 15 kG (1.5T). In extreme cases 17 kG or 18 kG is used if
absolutely necessary. In very large brushless machines where the length
of the return path becomes large, usable flux from the magnet is lost
driving it through the iron so lower densities should be used. If low
PWM chopping frequencies are used in the power section to control
voltage or limit the current, the eddy-current and hysteresis losses may
cause excessive heating in the core iron if the densities are too high.
Laminating the outside rotor shell is not practical as it is in interior
rotors, so the flux density should be kept at 10 or 12 kG for these
applications.
Allowance on magnet thickness for demagnetizationSome designers and
magnet application engineers suggest a magnet thickness to allow 10%
3-24
3. B a sic d e sig n c h o ic e s

demagnetization at the pole tips so that the minimum amount of magnet


is used. The author does not agree with this notion as being relevant to
any brushless motor application known. The magnet thickness should be
determined such that the maximum current which the electronic drive
can provide (with a given current limit setting) will not demagnetize the
rotor magnets. This principle does not necessarily apply to PM brush
motors used with battery power such as automotive cranking motors.
The author strongly encourages the application of this principle in the
design of all brushless DC motors.
Use of intrinsic demagrutkatiDn curveManufacturers of Ferrite and rare
earth products provide an intrinsic demagnetization curve as well as the
normal curve (Chapter 16). The normal curve is used to determine the
amount of flux produced by the magnet which links the stator phase
windings to produce torque. The intrinsic curve is used to determine the
maximum stator current allowable before magnet demagnetization occurs
(usually at the trailing-edge pole-tip). The demagnetizing effect of phase
current on the normal curve is shown in Fig. 4.4.
Maximum stator current (demag current)application of Amperes Law
around a magnetic flux line embracing the maximum possible ampere-
turns per pole per airgap gives the MMF which tends to demagnetize the
magnet. If HA is the maximum negative //which can be withstood in the
magnet, then the maximum line current in a three-phase wye-connected
motor must be limited to a value less than /rfcmag given by

4it * 39.37 J z/2p (3.6)


z
in [A], where and g are in inches and HA is in Oersteds; zis the total
number of conductors actually carrying current; p is the number of pole-
pairs; and a is the number of parallel paths in the winding. (See also
equations (4.27) and (4.28)). The term in square brackets is equal to
2.02. This formula is based on an Ampere's Law contour that encloses
z /2p conductors and goes through two magnets (Zj^) and two airgaps (g)
in series. With two phases in series z is given by

3-25
D e s ic n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

z = 2 * 2aNph = 4 ^ ph, (3.7)


remembering that one turn equals two conductors and iVph is the
number of turns in series per phase. An alternative expression for z in
a wye connection is
z = | x tumsjcoil x coils!phase * phases. (3.8)
where Z is the total number of conductors in the whole machine. With
a three-phase unipolar drive (Chapter 5), only one phase is on at any
time and the 4/3 factor in equation (3.8) should be replaced by 2/3.
With a two-phase motor, z = Z/2 and the 4/3 factor should be replaced
by 1. With a delta-connected motor, the 4/3 factor in equation (3.8)
should be replaced by 2, and /demag is the maximum permitted line
current. These relationships are summarized as follows:
Wye i= 2/3 x Z
Delta i- Z
Unipolar m
2-pha$c z* Z/1

T able 3.12 C o n d u c t o r s r ele v a n t t o c a l c u l a t io n o f d e m a g c u r r e n t

Use of intrinsic B /H curve for the correct temperatureIt must be mentioned


in connection with the calculation of the stator field reaction effects on
the magnet demagnetization that if NdFeB magnets are to be used,
special care must be exercised to use the intrinsic demagnetization curve
for the highest operating temperature of the magnet. The intrinsic
coercivity of these magnets is seriously degraded at elevated
temperature, whether they are bonded or sintered. Limiting the
operating temperature and/or adding thickness to the magnet can
usually yield the desired protection against demagnetization. If Ferrite
magnets are used, the intrinsic curve of the lowest operating temperature
must be used to determine the load line when exposed to the stator
MMF. Ferrite magnets have the unusual characteristic of a positive
temperature coefficient of intrinsic coercivity HQl (in the normal range
of applications up to 150 C), while the rare-earth magnets have a
negative temperature coefficient of Htci.

3-26
3. B a sic d e s ig n c h o ic e s

3.7 Lamination and stator stack design


Cutting the lammatumsThe stator of the brushless machine consists of a
Stack of soft iron laminations, an insulation system, copper phase coils,
and lead wires or connectors. For prototypes the laminations may be cut
by numerically-controlled laser cutting or wire-EDM (electro-discharge
machining), sometimes known as wire-slotting. Both these methods give
accurate dimensions (as good as 0.0005"), but the wire-EDM tends to
give a squarer edge. For production motors the laminations are cut using
a progressive die and an automatic press. For low volume production,
the progressive dies are made to punch a single set of laminations per
stroke (one stator punching and one rotor punching). The result is that
the material used is not the area of the lamination cross section, but the
square of a number slightly larger than the maximum lamination outside
diameter. If round lams are used, then about 21% of the material is
waste. If the volume is expected to be high, a quotation on a progressive
die which will punch three at a time is recommended. This will greatly
reduce the price of the laminations, pardy because the punch press
utilization is tripled, and partly because of material savings due to nesting
of three rows of punches in the die, using a strip which is not as wide as
three times the stator diameter.
Choice of can plateIn general, a high-performance motor (one with high
efficiency and/or high operating speed) will require thinner laminations
of materials that have low specific core losses. If the volume is low and
cost is not as important as performance, the use of M-19 of either #29
gage or #26 gage is recommended. Ultra-high speed motors designed for
high power density will require cobalt-iron laminations of very thin gage
(.005" or .010"). Most other brushless motors can use either #26 or #24
gage M-36 or M-43 laminations. However, for very high volume
competitive applications, the testing and evaluation of several gage
thicknesses, heat treatments, core plates, and materials should be
Conducted to determine the most cost effective lamination material.
Stacking-1The next consideration has to do with the actual stacking of the
laminations into a pack. There are essentially five ways to hold a
lamination stack together before winding for the stator:
1. Bonding with epoxy or Loctite
2. Riveting
3-27
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

3. Welding
4. Self-cleating
5. Slot liners
The bonding method is usually used only for prototypes or small
quantities. It is obvious by its name that epoxy in the form of a thin spray
coating is applied to at least one side of every lamination before they are
stacked in a fixture, clamped, and heated to cure the epoxy. Another
form of this is to align, clamp, preheat and apply Loctite which
penetrates between the laminations and bonds them together. Another
method is to provide holes in the stamping die to accept through-rivets
which are staked in place. The holes must be located at points in the
magnetic circuit such that the electrical circuits formed by shorting the
laminations to the rivets will not have significant induced currents in
them. The rivet holes are usually near the outside diameter of the
lamination. Rivets or overboils can be fitted outside the lamination stack.
In some cases they also help to hold the end-bells in position.
Another common procedure is automatic TIG (tungsten/inert gas)
welding. This can be used on an automatic assembly line for low-cost,
high-volume production.
A popular method of stacking and retaining laminations into packs that
do not require welding is to use a progressive punching die with a station
at the end of the stamping process which makes a small indentation in
each lamination. The indentations protrude on the other side of each
lamination, providing a self-cleating action when they are nested together
under pressure. The indentations are usually located in the
neighborhood of the teeth.
A common way to hold stacks together until they are wound during the
production of AC induction motors is easily used for brushless DC
motors because of the similarity of the stator laminations. This method
uses cuffs on each end of the slot liners which are automatically inserted
into the stator slot openings. The cuffs on either end of the lamination
stack are folded back automatically and hold the stack together quite
snugly until it is wound and varnished. This is an excellent method for
high speed automatic brushless motor manufacturing which can be
borrowed or adapted from induction motor production lines. Skewed
packs can be made after winding using this system.

3-28
3. Basic d e s ig n c h o ic e s
Insulating the slots1O nce the lamination stack is in a pack form, it must
be insulated before winding so that the magnet wire does not short out
to the pack. There are several ways that this is accomplished. For
prototypes either hand-cut slot-liners and end insulators are used, or if
the equipment is available a 3-M fluidized-bed epoxy coating system can
be used. For high-volume production motors a molded plastic insulator
is commonly used for small brushless machines, usually one on each side
of the pack. Frequently, in the molding process of the insulators various
connection methods are molded as part of the insulator so that
automatic connections of lead wires, Hall switches or printed circuit
boards can be incorporated with the attachment of the magnet wire from
(he coil windings. This high-volume insulating and connecting technique
is seldom used in large motors. The most common method for larger
motors is the cuffed slot liner as used in AC induction motors.
Automatic equipment is available for this sort of insulating.

fig. 3.3 Lamination packs with slot liners (courtesy of Industra Automation)

3-29
D esig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

A separate insulation system such as the molded plasdc pieces or the


cuffed slot liners is desirable for high-reliability insulation systems. Any
sort of epoxy coating technique is subject to difficulties with process
control, and hi-pot4 failures are strong possibilities. Fig. 3.3 shows several
lamination slacks with cuffed slot liners. The height of the cuff above
each side of the pack of laminations is slighdy higher than the minimum
specified by Underwriters Laboratories to assure that the cross-over of
the coil end-tums cannot be shorted on the laminations.
Winding and inserting the phase cmbThe phase coils can be placed into
the insulated lamination pack in several ways. The method used for
prototypes is to make a coil form of the approximate size, hand-wind the
coils, and hand-insert them into the correct slots. Usually the coils in a
set or phase are wound continuously, with the finish lead of the first coil
continuous with the start lead of the second coil and so on to minimize
the number of interconnections. For production manufacturing of
brushless motor stators, if the throw or the distance between the start
side of a coil and the finish side of a coil is not more than one or two
slots, needle winders can be used to wind the coils in place. It is also
possible to use a threc-needle machine with three starts and wind all
three phases simultaneously, indexing after each coil is completed to the
next coil rotation with continuous interconnections of magnet wire.
The other automatic method of winding is to use AC induction motor
winding equipment. Usually the coils of the phases are wound separately
and laid in slots of tooling which fits the particular lamination pack.
Either hand loading a stator pack or automatic loading of a stator pack
to a second station permits automatic insertion of the finished coils into
the insulated slots of the pack. This process is sometimes called "coil-
shooting". It is important to note that when hand insertion of the phase
coils is used the winding pattern can be of the type known as a "lap
winding". Needle winders can wind lap windings, but intricate tooling
is required and this method is usually limited to either single-slot-pitch
or two-slot-pitch windings. For automatic winding and coil insertion
using modem expensive machinery the only possible winding pattern is
known as "concentrated windings."

^ Hi-pot refers to the high-voltage withstand test applied to the insulation of


finished motors.
3-30
3. B a sic d e s ig n c h o ic e s
VarnishingThe end turns are usually formed and ded or laced before
varnishing the entire wound completed stator assembly. Varnishing or
encapsulating is necessary to provide a heat conduction path from the
inner copper turns to the lamination stack. The second reason for
vamish is to eliminate movement of adjacent turns as current is switched
on and off. Over a prolonged period of time if the conductors are not
held by varnish, the motion between them caused by electro-magnetic
forces would eventually wear through the varnish insulation on the wires
and cause electrical shorted turns. A third reason for varnishing is to
provide a higher dielectric strength than the air which would otherwise
lie between the turns of insulated magnet wire. Finally, the varnish
excludes moisture, dust and chemicals which could have a deleterious
effect on the integrity of the insulation.
Wiruimgunlh multiple stmnd conductorsMany times when the final number
of turns is calculated for a specific brushless design the size of the copper
conductor is too large to wind by one of these methods. It is not
uncommon to wind several strands of insulated magnet wire in hand
using finer gages to be equivalent to the cross section of the desired
gage. This, of course, provides the same phase resistance as the single
conductor, but allows a more flexible set of coils to be wound more
effectively. One of the problems with multiple-strand conductors is the
difficulty of attaching lead wires. It is difficult to strip and solder a
multiple- strand conductor to a heavy lead wire. The best way to make
this attachment for a good electrical joint without voltage drops, is to
fuse or weld the multiple strand magnet wire to the lead wire. Many
brushless servo motors use MS connectors for power leads and
commutation signal leads. By utilizing these heavy connectors for power
leads it is possible to eliminate the lead wire and bring the ends of the
magnet wire from the phase windings directly to the back of the MS
connectors and solder or weld them in place. There is a considerable
amount of controversy over which is the best method, but each one
should be carefully analyzed and the correct one selected for a given
manufacturing operation.
Minimization of lossesAt low speed the predominant loss is the P R or
copper loss. The way to minimize this is to fill the slots with as much
copper as possible: in other words, to use the maximum possible wire
gage. Cramming copper into the slots not only reduces the current
density, it also improves the thermal conduction of heat from the

3-31
D e s ig n o f b r u sh l e ss p e r m a n en t -m a g n e t m o t o r s

conductors to the lamination stack, especially when the coils are


varnished or encapsulated. Maximizing the wire gage reduces the phase
resistance R, but it has no significant effect on the inductance or the
back-EMF because these depend on the number of turns, not on the wire
gage-
The second largest loss component in brushless motors is the core loss,
which is due to eddy-currents and magnetic hysteresis in the laminations.
The eddy-current losses can be reduced by using thinner laminations, or
by using core plate with a high electrical resistivity such as 1.5% or 3%
Silicon steel. Both these alternatives increase cost. For very high speed
motors, the use of Silicon steel and thin laminations must be combined
with a low level of flux-density especially in the stator teeth. However, in
many cases high-speed motors are used in applications where direct or
indirect liquid cooling is available, permitting a very high loss density
without causing excessively high temperature in the core and windings.
The losses in core plate are discussed further in Chapters 9 and 16. They
are not amenable to accurate calculation, and are liable to vary as a
result of temperature, mechanical stress, and manufacturing
imperfections either in the heat treatment or in assembly. The non-
sinusoidal waveforms in brushless DC motors induce additional core loss
components that are of little significance in line-frequency AG motors.
Until recently, core loss data at different frequencies and flux densities
has been scarce, but many steel companies and research organizations
are making data available because of the widespread use of variable-speed
electronically controlled motors.
A fundamental limitation of brushless permanent-magnet machines is
that the field of the magnets cannot be varied. This means that the no-
load core losses at high speed have to be tolerated and cannot be
reduced as they can in a wound-field machine.
Number of stator teeth and core dimensionsOnce the number of stator teeth
has been determined and the arc of the rotor poles has been set, it is
necessary to determine the number of stator teeth which will have to
carry the flux from a single pole. This is done by comparing the rotor
pole span with the stator tooth span. If only one tooth spans a pole at
a time then the width of that tooth must have enough cross-section to
carry the total flux of the pole. If two teeth span one pole then obviously

3-32
3. B a sic d e s ig n c h o ic e s

the cross-sectional area of the teeth in the stator can be reduced


proportionately. As previously mentioned, the peak flux density in the
stator teeth and stator yoke should be kept below 15 kG. If extremely
high speed is expected then the flux density may need to be limited by
using larger cross-sections of iron to minimize core losses.
The thickness of the back iron or yoke of the stator must be determined.
Generally the minimum cross-section needs to be only one-half of the
sum of the cross-sections of the teeth under one pole, because the flux
is divided by two as it splits and goes both directions around the stator
yoke.
Stator tooth-tipsOne of the most important aspects of lamination design
for brushless motors using permanent magnets is the radial depth of the
tooth-tips ( ( i n Fig. 5.30). Many times these stator tooth-tips are
permitted to be too thin and are subject to saturation. As the edges of
the magnet poles rotate and overlap the protruding tooth-tips, saturation
occurs and cogging or detent torque results. This saturation of the tips
has a similar effect to having a very wide slot opening between the teeth.
As a rule of thumb, the thickness of the tooth-tips should be the same
dimension as their width, plus a generous radius in the corners so that
saturation does not occur. This point in lamination design is often
overlooked by the brushless motor designer and is a simple way to
reducing cogging torque. In some small motors used in disc drives where
noise is an important issue, the minimizing of cogging torques can have
a significant effect.
Cogging, skewIt has previously been stated that the use of integral-slot
designs causes extensive cogging or detent torque. An integral-slot design
is one in which the number of rotor poles divides evenly into the number
of stator slots. The cogging is exacerbated by the fact that every edge of
every pole lines up with stator slot openings as the rotor rotates. If the
number of rotor poles does not divide evenly into the number of stator
slots the cogging or detent torque is greatly reduced. The price paid for
using fractional-slot designs is a possible reduction in torque, caused by
a reduction in the winding factor (for example, C in equation (3.3))
resulting from restrictions on the distribution of phase coils with an odd
or non-integral number of slots/pole.
An effective way to reduce the cogging torque is to skew the stack of

3-33
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

laminations a quarter or a half of a slot-pitch. It is possible to wind the


stator on a loose lamination stack held in place by cuffed slot liners, skew
the stack after winding, and finally weld in three places on the OD to
hold the stack in the skewed position. Several manufacturers use this
method because it is easier to insert coils or wind with a straight stack.
Management of CTuf-*umsMany brushless stator configurations, especially
those evolved from AC induction motor products, exhibit large end-turns
from slot to slot on each end of the stack. In order to contain these end-
turns, it is necessary to lace them with string before varnishing. This can
be done by hand for small quantities or the use of automatic lacing
machines can be employed if they are available. The point is, that it is
necessary to retain these end turns before varnishing to ensure that their
location is controlled and does not result in any shorting situations
caused by wear of the varnish insulation. Before the end turns are laced
it is usually advisable to interleave sheets of woven fiberglass between the
end turns of the coils for lap windings which is trimmed before lacing.
Large end turns can be compacted in a process known as "blocking" or
"forming" in a set of blocking dies.

3.8 Winding Arrangements


The author is not familiar with every possible winding arrangement for
polyphase motors that engineers have devised over the past 150 years.
Only the most common and apparently practical ones will be discussed
here. The windings of brushless permanent-magnet motors with interior
rotors can be similar to those of three-phase induction motors, including
lap windings and concentric windings. In other cases the windings can
be similar to those of shaded-pole motors, with concentrated coils wound
around a single tooth. Exterior-rotor motors have stator windings that are
physically similar to those of DC motor armature windings except for the
terminal connections and the number of leads.
3.8.1 Coil span
A rule of thumb is that the maximum coil span or winding pitch omax
should be equal to the next lowest integer obtained after dividing the
number of rotor poles into the number of stator slots. This can be
formulated as

3-34
3. B a sic d e sig n c h o ic e s

= Max coil span (slots) - NLI Stator slots (jVslots) (3.9)


Rotor poles (2p)
where "NLI" means "next lowest integer" after performing the division.
For example, with 18 stator slots and an 8-pole rotor, 18 + 8 = 2-25. The
preferred coil span would be the maximum span of two slot-pitches. If
the same 18-slot stator were used with a 6-pole rotor, 6 poles would
divide into 18 slots 3 times exactly, so the maximum coil span would be
3. This is often expressed as "slots 1 and 4". Note that the preferred coil
span can be less than the maximum span (this is called "short-pitching"
or "chording"). Also, the coil span is not directly affected by the number
of phases.
3.8.2 Coils per pole
Windings should be distributed as evenly and uniformly as possible
around the circumference of the stator for each phase. This
arrangement will yield the highest utilization of active conductors
producing torque at all rotor positions.
The number of coils per pole is an important factor in the winding
design. Although special computer programs are available to design
winding patterns for any number of slot and pole combinations (Chapter
12), it is always desirable to use winding diagrams to define the
distribution of phase conductors. Fig. 3.4 shows an example of a four-
pole rotor with four different stators, with their respective locations of
coils. With the 6-slot stator the coil span is 1 slot-pitch, in agreement with
equation (3.9), and with three phases there can be essentially only two
coils per phase (unless the machine has dual windings). Note that there
is no overlap between any of the coils in this machine. This is not only
a distinct manufacturing advantage, but it also reduces the likelihood of
an interphase fault. The end-turns are veiy short in this winding.
With the 12-slot stator in Fig. 3.4, the winding is shown with a coil-span
of 3 slot-pitches. This permits 4 coils per phase or one coil per pole in
a double-layer winding (as shown). Alternatively, if every other coil is
omitted, there would be 2 coils per phase and the winding would be a
consequent-pole single-layer winding. With three phases, neither of these
winding arrangements has any slot containing conductors from two
different phases. A variant suited for sinewave motors is to use a coil-span
3-35
D esig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

of 2 slots, which produces a more sinusoidal back-EMF and shorter cnd-


turns; however, with three phases, all slots would then contain conductors
from two different phases.

1/2 COIL/POLE/PHASE 1 COIL/POLE/PHASE

6 SLOT 12 SLOT
4 POLE 4 POLE

1 COIL/POLE/PHASE 2 CO IL/POLE/PHASE

24 SLOT 36 SLOT
4 POLE 4 POLE

Fig. 3.4 Determination of conductors per ptuue per pole

3-36
3. B a sic d e s ig n c h o ic e s
With the 24-slot stator in Fig. 3.4, the coil span could be 6 slots, but with
1 coil per pole and three phases, half the slots would be empty while the
other half would contain two coil-sides each. Fig. 3.4 shows a more
practical arrangement with a coil span of 5 slots and one coil per pole.
With the 36-slot stator the coil span could be 9 slots but 8,7,6 or even 5
could be used. The winding shown has 2 coils/pole with spans of 9 and
8 each. Alternatively the spans could be 8 and 7 slots. This is a concentric
winding. In an electrically equivalent lap winding, all coils would have the
same span, either 8 or 7, and there would still be two coils per pole.
The windings in Fig. 3.4 are all relatively concentrated, as would be
suitable for squarewave motors. For sinewave motors it is more
appropriate to use short-pitched coils and a more distributed winding.
The windings in Fig. 3.4 fill exactly 1/3 of the available space in the slots,
leaving the same amount of space for the other two windings.
4 POLE ROTOR 6 COILS/POLE

ST ARTS FINISHES

Fig. S.5a Lap winding

3-37
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

4 POLE ROTOR 6 COILS/POLE

A B C A B C

I I
I
T
1
4
A B C A B C
START S FI NI SHES
Fig. 3.5b Concentric winding
An example of a lap winding is shown in Fig. 3.5a, and an example of a
concentric winding in Fig. 3.5b.
3.8.3 Winding configurationsfractional and integral slot
An integral-slot stator is one with an integral number of slots per pole, so
that the pole-pitch is an integral number of slot-pitches. The windings of
integral-slot stators are naturally regular and symmetrical: the coils fall
naturally into groups, each group usually being associated with one pole
as in the 36-slot stator in Fig. 3.4.
A fiactional-slot stator has a non-integral number of slots per pole: for
example, a 4-pole motor with 15 slots has 3.75 slots/pole and there is no
perfectly symmetrical winding arrangement with identical groups of coils

3-38
3. B a sic d e sig n c h o ic e s

associated with single poles. A three-phase motor with this slot/pole


combination is shown in Fig. 3.6. Building a fractional-slot winding
arrangement provides scope for considerable ingenuity and there is
probably no single procedure that results in an optimum winding for all
slot/pole combinations. A simple procedure which gives good general
results for double-layer windings is described below.

Fig. 3.6 Three-pha3e 15/4 motor fractional-slot winding

The first step is to calculate the slots/pole and the maximum coil span
ffraax us*ng equation (3.9). 0max is an integral number of slot-pitches,
and the remainder e is a fractional number of slot-pitches between 0 and
1 : thus
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

For the 15/4 motor, <Jmax = 3 and e = 0.75. The actual coil span used
can be less than 3 slots, and in the example shown in Fig. 3.6 the coil
span is a = 2 slots.
The first coil in Fig. 3.6 is wound in slots 1 & 3, so the position of the
axis of this coil can be charactemed as (1 + 3)/2 = 2, i.e. 2 slot-pitches
from the origin (the w-axis), or aligned with the centreline of slot 2.
The rule for locating subsequent coils is expressed in terms of Sp, the
number of slots forward from the "return" coilside of the previous coil to
the "go" coilside of the coil next to be inserted. The value of is given
by
=
f CTm a x + 1 if e S 0 .5
(3-11)
I ^ s s -t7max> f > 0-5
where Na is the number of slots per section. The winding is divided into
sections if the number of coils per phase Cph is divisible by the number
of pole-pairs p. In fact, the number of sections S is equal to the highest
common factor of and pole-pairs p, so that
N = -----^ 2 * -----. (3.12)
H C F [^h^ ]
The rule expressed in equations (3.11) and (3.12) tends to minimize the
build-up of phase displacement or phase error between subsequent coils.
Note that when Sp = cj nax + 1 the winding is progressive, in the sense that
successive coils follow one another in the positive (CCW) direction.
When Sp = Na - omax, the winding is retrogressive in the sense that
successive coils follow one another in the negative (CW) direction.
The winding in Fig. 3.6 has only one section: the number of coils per
phase is 15/3 = 5, and this has no common factor with p, which
is 2. Therefore Nss = Njloti = 15, and with CTmax = 3 and e = 0.75 > 0.5, .Sp
= 3 + 1 = 4 slots. Since the "return" coilside of coil 1 is in slot 3, the "go"
coilside of coil 2 will be in slot 3 + 4 = 7 . The polarity of successive coils
is alternated so that coil 2 has a span of -2 rather than +2, and its
"return" coilside is therefore in slot 7 - 2 = 5 . Note that the axis of coil
2 is at (7 + 5)/2 = 6, and this is 4 slots further on than the axis of coil 1.
Since the pole-pitch is 3.75 slots, the axis of coil 2 is displaced 1/4 of a

3-40
3. B a sic d e s ic n c h o ic e s

slot clockwise from the position it would need to have in order for the
EMF's in coils 1 and 2 to be in phase. It can be said that the phase
displacement between coil 2 and coil 1 is + 1 /4 slot. This does not sound
much, but in clectrical degrees 1 slot is equivalent to 1 /1 5 x 2 x 360 =
48, so the phase displacement of coil 2 is 12 relative to coil 1 .
Proceeding with the winding, coil 3 is in slots (5 + 4) = 9 and (9 + 2) =
11. Its axis is at (9 + l l ) / 2 = 10, that is, 8 slot-pitches further on than the
axis of coil 1. The nearest integral number of pole-pitches from the axis
of phase 1 is at 7.5 slot-pitches from the axis of phase 1, and therefore
coil 3 has a phase displacement of 1 /2 slot or 24 relative to coil 1.
The winding can be completed by continuing in the same fashion: coil
4 is in slots 15 & 13 with a phase displacement of 36, and coil 5 is in
slots 2 & 4 with a phase displacement of 48 or one complete slot. The
resulting coil list is summarized in Table 3.13.
Coil No. Turn* Go slot Retn dot Span
1 12 1 3 2
2 12 7 5 -2
3 12 9 11 2
4 12 15 13 -2
5 12 2 4 2

T a ble 3.13 C o il L is t f o r 3-ph ase m o t o r w it h 15 sl o t s a n d 4 po le s

H ie phase displacement accumulates in a regular fashion from one coil


to the next, producing the same effect as in a distributed lap winding.
There is therefore no difficulty developing appropriate winding factors
for windings built by this procedure: see Chapter 6.
When there are multiple sections, the winding algorithm expressed in
equations (3.11) and (3.12) is applied only until one sequence of coils is
completed. The first sequence is the first Cph/S coils generated by the
winding algorithm, and these do not necessarily belong in the same
section. A sequence is a logical grouping of coils, whereas a section is a
physical sector of the stator. An example should make this more clear.

3-41
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

Big. 8.7 Three-phaje 18/8 motor fractional-slot winding

The three-phase motor in Fig. 3.7 has 18 slots and 8 poles, and C. , =
18/3 = 6 coils per phase. With p = 4, the highest common factor o f fph
and p is 2, and therefore there will be two sections and two sequences,
each containing 3 coils. Each section will have 4 poles and = 9 slots.
The number of slots/pole is 18/8 = 2.25, so oraax = 2 and e = 0.25.
Therefore, from equation (3.11), 5^ = 9 - 2 * 7. These parameters
define the winding pattern shown in Fig. 3.7 for the first three coils only,
forming the first sequence. The second sequence is obtained by copying
the first sequence coilside-for-coilside, Na slots further on. Thus coil 1 in
slots 1 Sc 3 is followed by coil 2 in slots (3 + 7) = 10 and (10 - 2) = 8,
then the first sequence is completed by coil 3 in slots (8 + 7) = 15 and
15 + 2 = 17. The second sequence is obtained by copying coil 1 to coil

342
3. B a sic d e sic n c h o ic e s

4, 9 slots further on in slots (1 +9) =10 and (3 + 9) = 12. Coil 5 is a


copy of coil 2 in slots (10 + 9) = 19 mod 18 = 1 and (8 + 9) = 17, and
coil 6 is a copy of coil 3 in slots (15 + 9) = 24 mod 18 = 6 and (17 + 9)
26 mod 18 = 8. This naturally produces a winding in two sections, the
first containing the odd-numbered coils 1,3 and 5 and the second the
even-numbered coils 2,4 and 6. The resulting coil list is summarized in
Table 3.14, with the two coil sequences separated by a double line.
Coil No Turns Go slot Retn slot Span
1 12 1 3 2
2 12 10 8 -2
3 12 15 17 2
4 12 10 12 2
5 12 1 17 -2
fi 12 6 8 2

T a b le 3.14 C o il L ist fo r 3-ph ase m o t o r w ith 18 sl o t s a n d 8 p o l e s


The axis of coil 1 is at slot (1 + 3)/2 = 2 and the axis of coil 2 is at slot
(10 + 8)/2 = 9, i.e. displaced from the axis of coil 1 by 7 slots. With 2.25
slots/pole the nearest integer number of pole-pitches subtending an arc
close to 7 slots is 3, i.e. 3 x 2-25 = 6.75 slots. Therefore the phase
displacement between coil 2 and coil 1 is 0.25 slots or 0.25/18 x 4 x 360
= 20 electrical. With reference to equation (6.36), the effective
fundamental distribution factor of this winding is =sin (3 x 20/2)/3
sin (20/2) = 0.960. The 5th-harmonic distribution factor is AdS = sin (5
x 3 x 20/2)/3 sin (5 x 20/2) = 0.218. The fundamental pitch factor is
Apl = cos (e/2) and with e = 0.25 slots = 20 elec, Apl = 0.985. Similarly
fcp5 = cos (5 x 20/2) = 0.643. Thus the overall winding factors are ^
= Adl&pl = 0.946 for the fundamental and = Ad5& g 0.140 for the 5th
harmonic, suggesting the possibility of a fairly good sinewave back-EMF.
Higher-order harmonic winding factors can be found in similar manner.
Fig. 3.8a shows the back-EMF waveform from a single coil of the winding
in Fig. 3.6, and Fig. 3.8* shows the line-line back-EMF waveform from the
whole winding connected in wye.

3-43
D e s ig n o f b r u sh l e s s pe r m a n en t -m a g n e t m o t o r s

Fig. 3,8 Back-EMF waveform for three-phase 15/4 motor of Fig, 3.6.
a EMF waveform of one coil
b line-line EMF waveform of entire winding {wye connection)
c Electromagnetic torque waveform with squarewave drive
Also shown in Fig. 3.8 is the electromagnetic torque waveform obtained
with perfect squarewave drive. Similarly, Fig. 3.9a shows the back-EMF
waveform from a single coil of the winding in Fig. 3.7, and Fig. 3.9b
shows the line-line back-EMF waveform from the whole winding.

3-44
3. B a sic d e s ig n c h o ic e s

Fig. 3.9 Back-EMF waveform for three-phase 18/8 motor of Fig. 3.7.
a EMF waveform of one coil
b line-line EMF waveform of entire winding (wye connection)
c Electromagnetic torque waveform with squarewave drive

The winding is connected in wye. Fig. 3.9 c shows the torque waveform for
squarewave drive. In both cases the magnet pole arc is 140 electrical,
and the single-coil waveforms were calculated w i t h 5 x ! 2 = 60 turns in

3-45
D e s ig n o f b r u sh l e s s pe r m a n en t -m a g n e t m o t o r s

Fig. 3.8 and 6 x 12 = 72 turns in Fig. 3.9. The single-coil waveforms


appear as "phase" EMFs, and the 30 phase-shift between the phase EMF
and the line-line EMF is apparent in these figures. In both motors the
EMF of the whole winding is much closer to a pure sinewave than that
of a single coil.
For integral-slot windings the winding algorithm expressed by equations
(3.11) and (3.12) works equally well as for fractional-slot windings, the
number of sections being generally equal to the number of pole-pairs.
Wharr. to start winding the second and third phasesIn a three-phase motor
the axes of the phase windings should be displaced from one another by
120 electrical degrees. If phase 1 starts in slot 1, the start of phase 2
should be in slot 1 + Offset, where Offset is the number of slot-pitches in
2/3 of a pole-pitch or 2/3 x Nilo{s/2p. If this number is not an integer,
it is necessary to search for alternative starts using the formula

Offset = 1 x + k= i i2 3 (3.13)
3 2p p
With integer values of k, the second term merely advances the start of
phase 2 by 360A electrical degrees. Some examples of windings are given
in Figs. 3.113.19, and the Offset parameter is tabled in these figures.
For example, in Fig. 3.11, with 6 slots and 2 poles {p = 1), equation
(3.13) gives Offset = 2 with no need for the second term, i.e. k = 0. In Fig.
3.12, with 9 slots and 8 poles, 2/3 x 9/8 = 3/4, which is not an integer.
Trying k = 1 gives 3/4 + 9/4 = 12/4 = 3, which is the correct value of
Offset for this winding. If no integral value of Offset can be found from
equation (3.13), it is impossible to wind a balanced three-phase winding
in the given number of slots and poles. For example, with 15 slots and
6 poles Offset is nonintegral for all integer values of k. On the other
hand, the 15/4 motor works with k= 1 to give Offset = 10, as in Fig. 3.16.
Cheeking the windingA simple check on the validity of a winding
configuration generated by equations (3.11-13) is to test whether all the
slots contain equal numbers of coilsides. Generally this will be the case,
provided that valid slot/pole combinations are used (Tables 3.3-9).

3-46
3. B a sic d e s ig n c h o ic e s

Fig. 3.10 Back-EMF of one phase, commutated by squarewave drive

3.8.4 Examples of windings and back-EMF waveforms


The windings in Fig. 3.4 are not intended to represent optimal practical
windings, but merely to show the factors involved in choosing the coil
span and the number of coils per pole. The final choice should be made
only after calculating the back-EMF waveform and the winding factors
(see Chapters 5,6, and 8). These calculations are potentially time-
consuming and it helps to have guidelines based on the classification
according to the number of slots per pole as in Table 3.8.
3-47
D esig n o f bru sh less perm a nent -m a gnet m o t o r s
In Table 3.8 nine basic groups of motors were listed according to the
number of slots/pole, with the most common examples listed in each
group. Using a computer to synthesize the back-EMF waveshape, the
nine motor groups have been analyzed using an example of each. Fig.
3.10 shows a typical back-EMF waveform from a three-phase brushless
motor. In this example the waveform is very nearly sinusoidal.
The lower graph in Fig. 3.10 shows the effect of commutating the back-
EMF waveforms of all three phases with a squarewave drive. The
commutations occur every 60 electrical degrees, giving an interval of 60s
during which a squarewave drive would force a constant current against
the back-EMF. With constant-current drive, the resulting torque waveform
would follow the shape of the back-EMF waveform, and therefore this
diagram gives an idea of the inherent electromagnetic torque ripple that
would be expected with squarewave drive and perfect commutation of
the phases. The importance of the back-EMF wave shape for minimum
torque ripple with squarewave drive is clearly seen from this example.
Of course the sinusoidal back-EMF waveform with sinewave drive produces
no torque ripple in theory, but for squarewave drive it is useful to extend
the exercise depicted in Fig. 3.10 to all the basic groups of motor listed
in Table 3.8, with different coil spans and winding arrangements.
Example waveforms are given in Figs. 3.11-3.19. This data shows the
importance of the pole arc angle, the winding distribution, whether delta
or wye connections are used, and whether skew is used.
The example given in Fig. 3.11o shows four basic conditions, i.e.
wye/delta connections, with and without skew (of one slot-pitch). The
magnet pole arc is 2/3 of a pole-pitch. Figure 3.116 shows the same four
conditions, but with a magnet pole arc of one pole-pitch or 45
mechanical degrees (180 electrical degrees). The skew can be
accomplished by skewing the magnet poles or skewing the lamination
stack, or a combinadon of both. The back EMF waveshape would be the
same.
The variety of back-EMF wave shapes is quite extensive and indicates that
certain configurations are better than others for squarewave drive, while
others are suitable for sinewave drive. Figs. 3.11 through 3.19 show
examples from each of the motor groups in Table 3.8. The back-EMF
wave shapes (plotted on a basis of electrical degrees) are essentially the

348
3. B a sic d e s ig n c h o ic e s

same for every motor configuration within the same group, so the
examples shown can be applied to other motors in the same group.
The back-EMF waveforms in Figs. 3.11-3.19 were calculated without any
fringing effects, i.e. assuming no airgap .5 When actual back-EMF
waveshapes are measured on motors of these examples, fringing causes
the comers to be rounded: see Chapter 8. This does not significandy
affect the usefulness of this data in making comparisons between the
different configurations.
Some combinations of the windings and pole-arcs give rise to excessive
third-harmonic EMFs in the phases, making them unsuitable for delta
connection. This is discussed in more detail in Chapter 5.
All the plots in Figs. 3.11-3.19 extend over 180 electrical degrees. The
data with each curve includes the average value of the back-EMF
throughout the 60 commutation zone, expressed as a percentage of the
peak value of the ideal squarewave back-EMF that would be achieved with
a concentrated full-pitch coil (see Chapters 5,7 and 8). This percentage
expresses the effect of the coil span, skew, and method of connection on
the torque constant. The ideal commutation zone average would be
100%, but the actual value is reduced by short-pitching the windings, by
fringing at the edges of the magnet, and by skew. The reduction in
average EMF caused by skewing is noticeable and is the price paid for
low cogging torque. Other techniques for reducing cogging (radii or
chamfers on the magnets) have similar effects on the back-EMF wave.
The number of slots can be determined using the examples as a general
guide, along with the rotor pole arc angle to achieve the desired output
considering cogging torque and winding designs. In general, the smallest
number of slots gives the lowest labor cost in winding, and a coil span of
1 or 2 slots yields the lowest phase resistance. The phase inductance is
decreased if the same number of turns is distributed among a greater
number of slots, so the electrical time constant can be lower with a
greater number of slots per pole.

c The computer program, which is the work of G. Aha and R.C. Perrinc, constructs
the waveforms from a 35-term Fourier series. The truncation is the cause of the high-
frequency ripple visible on some of the waveforms (Cihb phrrwmrrwn).
3-49
D e sig n o f bru sh less perm a nent -m a g n et m o t o r s

Slots/pole 0.75
Slots 6

Poles 8

Phases 3
Pole arc 30 (120e)
Offset 2

Coil span 1

Hg. 3.11a

WYE DELTA
Skew

Comm, zone avg. = 100%


Skew = 1

Comm, zone avg. = 32.8% Comm, zone avg. = 37.5%

3-50
3. B a sic d e sig n c h o ic e s

Slots/pole 0.75
Slots 6
Poles 8

Phases 3
Pole arc 45 (180e)
Offset 2

Coil span 1

Fig. 3.11 fr

WYE DELTA
Skew

Comm, zone avg.


>100% Comm, zone avg - 100%
Skew = 1

Comm, zone avg. =37.5% Comm, zone avg. 43.7%

3-51
D esig n o f bru sh less perm a nent -m a gnet m o t o r s

Slots/pole 1.125
Slots 9
Poles 8

Phases 3
Pole arc 30 (120 e)
Offset 3
Coil span 1

Fig. 3.12a

WYE DELTA
Skew
Skew = 1

3-52
3. B a sic d e s ig n c h o ic e s

Slots/pole 1.125
Slots 9
Poles 8

Phases 3
Pole arc 45 (180 e)
Offset 3
Coil span 1

Fig. 3.124

WYE DELTA

Excessive
third-harmonic
Skew

Excessive
Skew = 1

third-harmonic

3-53
D esig n o f bru sh less perm anent -m a g n et m o t o r s

Slots/pole 1.5
Slots 6
Poles 4
Phases 3
Pole arc 60 (120e)
Offset 1

Coil span 1

Fig. 3.13a

WYE DELTA
Skew

Comm, zone avg. > Comm, zone avg. = 100%


Skew = 1

Comm, zone avg. = fi5.fi% Comm, zone avg. = 75%

3-54
3. B a sic d e s ig n c h o ic e s

Slots/pole 1.5
Slots 6

Poles 4
Phases 3
Pole arc 90 (180e)
Offset 1

Coil span 1

Fig. 3. ISA

WYE DELTA
Skew

Comm, zone avg. = 1(10%


Skew = 1

Comm, zone avg. = 75%

3-55
D e sic n o f bru sh less perm a nent -m a gnet m o t o r s

Slots/pole 2.25
Slots 18
Poles 8

Phases 3
Pole arc 30 (120e)
Offset 6

Coil span 2

Fig. 3.14a

WYE DELTA
Skew
Skew = 1

3-56
3. B a sic d e s ig n c h o ic e s

Slots/pole 2.25
Slots 18
Poles 8

Phases 3
Pole arc 45 (180e)
Offset 6

Coil span 2

Fig. 3.14ft

WYE DELTA

o Excessive
ii
third-harmonic
Skew
Skew = 1

Excessive
third-harmonic

Coram. zone avg. -92%

3-57
D esig n o f bru sh less perm anent -m a gnet m o t o r s

Slots/pole 3.0
Slots 12
Poles
Phases 3
Pole arc 60 ( 120 e)
Offset
Coil span
Fig. 3.15a

WYE DELTA

Comm, zone avg. - 100% Comm, zone avg. = 100%

Comm, zone avg. = 87.5% Comm, zone avg. * 100%

-58
3. B a sic d e sig n c h o ic e s

Slots/pole 3.0
Slots 12
Poles 4
Phases 3
Pole arc 90 (180e)
Offset 2
Coil span 3
Fig. 3.15*

WYE DELTA

o Excessive
li third-harmonic
Skew

Comm, zone avg. = 100%

Excessive
Skew = 1

third-harmonic

3-59
D esig n o f brushless perm anent -m a g n et m o t o r s

Slots/pole 3.75
Slots 15
Poles 4
Phases 3
Pole arc 60 ( 120e)
Offset 10
Coil span 3
Kg. S.lfia

WYE DELTA

J
/ \
Skew

Coram. zone avg -83%


V
Skew = 1

Comm, zone avg. - 80.5%

3-60
3. B asic CHOICES

Slots/pole 3.75
Slots 15
Poles 4
Phases 3
Pole arc 90 (180e)
Offset 10

Coil span 3
fig.

WYE DELTA

Excessive
third-harmonic

Comm, zone avg.


>96%

Excessive
third-harmonic
D esig n o f bru sh less perm a nent -m a gnet m o t o r s

Slots/pole 4.5
Slots 18
Poles 4
Phases 3
Pole arc 60 ( 120e)
Offset 3
Coil span 4
Fig. 3.17o

WYE DELTA
Skew
Skew = 1

3-62
3. BASIC DESIGN CHOICES

Slots/pole 4.5
Slots 18
Poles 4
Phases 3
Pole arc 90 (180 e)
Offset 3
Coil span 4
Fig. 3.17*

WYE DELTA
Skew

Excessive
Skew = 1

third-harmonic

SS3
D esig n o f bru sh less perm a nent -m a gnet m o t o r s

Slots/pole 5.25
Slots 21
Poles 4
Phases 3
Pole arc 60 (120 e)
Offset 14
Coil span 5
Fig. 3.18a

WYE DELTA
Skew
Skew = 1

3-64
3. B a sic d e s ig n c h o ic e s

Slots/pole 5.25
Slots 21

Poles 4
Phases 3
Pole arc 90 (180e)
Offset 14
Coil span 5
Fig. 3.1 BA

WYE DELTA


n Excessive
third-harmonic
Skew

Comm, zone avg. = 100%

Excessive
Skew = 1

third-harmonic

S-65
D esig n o f bru sh less perm anent -m a g n et m o t o r s

Slots/pole 6.0
Slots 24
Poles 4
Phases 3
Pole arc 60 ( 120e)
Offset 4
Coil span 6

Fig. 3.19a

DELTA

h
Skew

Coram. zone avg. = 87.5%


r 1
Comm, zone avg = 100%
L.

/
/ \
\'
Skew = 1

'

Comm, zone avg. - 87.5%


t
Comm. 2onc avg - 100%
\
3-66
3. B a sic d e s ig n c h o ic e s

Slots/pole 6.0

Slots 24
Poles 4
Phases 3
Pole arc 90 (180 e)
Offset 4
Coil span 6

Fig. 3.19ft

WYE DELTA

Excessive
third-harmonic

Comm, zone avg. = 100%

Excessive
third-harmonic

Comm, zone avg. = 100%

3-67
D esig n o f bru sh less perm a nent -m a gnet m o t o r s

3.8.5 Wire size calculations


The first step is to determine the total available winding space from the
stator lamination cross-section. The number of turns per coil has been
determined so that the maximum wire diameter (over varnish insulation)
can be calculated. An estimate of the maximum wire size is often based
on the permissible slot fill factor, If the winding pattern is a double
layer type with two coil sides per slot, a slot fill factor of 0.3-0.35 is
reasonable. If the winding is a single-layer winding with only one coil side
per slot, a slot fill factor of 0.65-0.7 can be used..
These slot-fill factors assume that the wire is square and perfectly nested,
i.e. with no gaps between the strands. For a given net slot area Aslot,
equal to the actual slot area minus the area taken up by the slot liner, if
the number of conductors per slot is Nthen the maximum wire diameter
D (measured over the insulation) can be determined from the equation

Aw ~ \ ^slot^slot
N
(3.14)

The next smaller wire gage is selected from a wire table, after selecting
the type of varnish insulation required for the required temperature
rating.
[NOTE: The definition of slot fill used here is not the same as the ratio
of conductor area to total slot area, which is considerably lower.]
3.8.6 Basic winding calculations
ResistanceThe actual "bare copper" wire diameter is used to calculate the
phase resistance. This requires the determination of the mean length of
one turn (MLT). The M LT includes twice the stack length, plus twice
the end-tum crossover length, plus twice the coil "bundle" thickness. A
full scale drawing or CAD print is useful for determining the MLT. For
concentric or concentrated windings, each coil in a pole group has a
different turn length, and this must be taken into account.
The resistance per coil Rc is a simple calculation using Q, the value of
resistance per 1000 ft. for the particular wire gage selected. Divide this

3-68
3. B a sic d e s ig n c h o ic e s

value by 12000 if the MTL is in inches. If Ne is the number of turns/coil,

R' = M LT x x Nr Ohms. (3.15)


^ 12000 c

The resistance per phase is determined by adding the Rc together, taking


account of series and parallel connections of the coil groups. If a
squarewave three-phase drive is used and the connection is wye, the
terminal or line-line resistance is twice the phase resistance. If the
connection is delta, the terminal resistance is 2/3 the phase resistance.
InductanceW hen the phases are commutated to the supply voltage as in
a squarewave drive, the build-up of current lags behind the voltage
because of the winding inductance. The electrical time constant (time
for current build-up) is equal to the phase inductance divided by the
phase resistance (Chapter 14). If this time constant is comparable to the
time taken to rotate 60 electrical degrees at a certain speed, the average
current during each commutation interval may be limited by the
inductance rather than the back-EMF. This effect is common in 12-volt
motors used in automotive applications and computer disc drives, where
it causes a concave droop in the speed/torque curve.
The self-inductance of a phase winding is considered to include several
terms due to the notionally separate components of flux established by
the phase current. These include flux that crosses the airgap, slot-leakage
flux, and end-tum leakage flux. Formulas for these inductance
components are given in Chapter 5. The mutual inductance between
phases is treated in a similar way. Both the self and mutual inductances
depend on the winding distribution in a way that makes hand-calculation
difficult, but computer methods (including the finite-element method)
are increasingly used: see Chapter 5.
3.9 Examples of motor construction
Fig.3.20 shows a classical design of brushless DC servomotor with 4 or
8 poles.The magnets are in the form of narrow bars held on to the rotor
by adhesive and glassfibre roving. The rotor core is laminated and a
brushless tacho-generator and position sensor are integrated in the
aluminum housing.

3-69
D esig n o f bru sh less perm a nent -m a gnet m o t o r s

Fig. 3.20 Bnuhlesi DC ervomotor (Bosch). Drawingsupplird by Univ.-Pmf. Dr.-fng.


G. Hmnebergrr of Institut f i r Elektrvche Maschmm, RWl'H-Aachm, [2]

Fig. 3.21 Brushlew DC disc motor (Infranor). Drawing mpplitd by Uhw.-Pnf. Dr-
Ing. 0. Hmntbrrgrr of Institut fur Ehtktritchr. Maschmen, RWl'H-Aachm, [2]

3-70
D e sig n o f bru sh less perm a nent -m a g n et m o t o r s

Fig. 3-21 shows a disc motor with two stator halves each carrying a set of
pole windings. The magnets are held in the rotating disc between the
$tator halves. The stator core is made from a wound-up strip of steel
bolted to the housing for good heat transfer. This motor has low inertia.

Fig. 3,22 Exterior-rotor 40mm brushleu DC motor for computer disc drive. Phots
kindly supplied by Syruktron Corp., Portland, Ongpn

Figs. 3.22 and 3.23 show typical exterior-rotor motors of the type used in
computer disc drives and laser scanners.

3-71
D e sig n o f bru sh less perm a nent -m a g n et m o t o r s

Fig. 3.23 Exterior-rotor half-height brushless DC motor for computer disc drive.
Photo kindly supplied by Synfktron Corp., Portland, Ortgon

Fig. 3.24 shows a small brushless DC actuator motor with a ballscrew


integral with the motor shaft.
Fig. 3.25 shows an 8-pole spoke-type rotor with circumferentially-
magnetized Ferrite magnets. This motor has excellent torque linearity,
low cogging (<1 %), a high resistance to demagnetization, and benefits
from the cost-effectiveness of using Ferrite magnets with their good
temperature coefficients and resistance to corrosion. Because of the flux-
focusing geometry, the air-gap flux-density is of the order of 0.9T, even
using Ferrite magnets. The pole-pieces are shaped to improve the back-
EMF waveform quality and torque linearity with sinewave drive.

3-72
3. B a sic d e s ig n c h o ic e s

Fig. 3.24 Small bruihleu DC actuator motor. Photo kindt) supplied by Lucas Aeroipacr

Fig. 3.25 8-pole spoke-typc rotor (Pacific Scicntific F40 Series). Photo kindly supplied
by Pacific Scientific, Rockford, lUinoii

3-73
D e sig n o f b r u sh l e ss pe r m a n e n t - m a g n e t m o t o r s

Fig. 3.26 Pacific Scientific brushless servomotors. The largest of these motors (7,5')
produce up to 50 Nm continuous stall torque and 200 Nm peak torque.
Photo kindly supplied by Pacific Scientific, Radford, Illinois

Fig. 3.26 shows a type of square-frame brushless servo motor popular in


a wide range of machine tools and motion control applications.
References
[1] Kenjo T and Nagamori [1985] Ptrmanml-magfut and brushkxi DC motors, Oxford
University Press.
[2] Ilenneberger G [1993J driven : a status mjiew. PCIM Conference proceedings,
Intelligent Motion, Nurnberg 1993, 1-14

3-74
4. MAGNETIC DESIGN
4.1 Introduction
The flux in a brushless permanent-magnet motor is established by the
magnets. We have seen in Chapter 1 that the torque is proportional to
the current and the flux, while the no-load speed is proportional to the
voltage and inversely proportional to the flux. The flux is clearly a most
important parameter in the design. This chaptcr describes the simplest
methods of calculating it, for later use in the calculation of the EMF and
torque equations and the speed/torque characteristic.

Fig. 4.1 CroM-iection of 2-polc motor showing the magnet flux

The simplest motor is the 2-pole motor, Fig. 4.1. The flux is intended to
link the coils of the phase windings on the stator, and these coils are
located as close as possible to the magnet to minimize the amount of
magnet flux that "leaks" from the N pole to the S pole without linking
any turns of the windings. The laminated steel core of the stator acts as

4 rl
D e sig n o f bru sh less perm anent -m a g n et m o t o r s

a flux guide. The high permeability steel teeth draw the flux radially
across the narrow airgap and the yoke (back-iron) returns it from the N
pole to the S pole with very little expenditure of MMF (magnetomotive
force or magnetic potential drop). The rotor hub performs a similar
function inside the rotor. Because the steel stator and rotor absorb very
little MMF, most of the magnets MMF is available to drive flux across the
airgap.
The slotting is an ingenious way to achieve a narrow airgap length while
keeping the winding conductors close to the magnet. Other benefits of
slotting are subtle but extremely important. The slotted structure
provides a rigid housing for the windings and their relatively fragile
insulation. It greatly increases the surface contact area between the
windings and the steel, providing a path of low thermal resistance which
is important in keeping the windings (and hence also the magnets) cool.
The steel does not have unlimited capacity for carrying flux. If the flux-
density exceeds approximately 1.6-1.7 Tesla [TJ, the permeability
decreases rapidly. At a flux-density of about 2.1 T, the incremental
permeability of steel is practically the same as that of air. The magnetic
design should ensure that the flux-densities in the steel are kept below
these levels, otherwise the magnet MMF will be wasted in driving flux
through the steel. The end result would be that the flux linking the
windings would be restricted; or, what amounts to the same thing, the
amount of magnet material required to establish a given flux would be
greatly increased.
Another reason for limiting the flux-density in the steel, especially in the
stator, is that the core losses increase rapidly at high flux-density. Core
losses are caused by hysteresis and by eddy-currents (see Chapter 9). The
eddy-current component can be reduced by stamping the laminations
from thinner-gauge sheet, or by using high-Silicon steels, but both of
these measures increase cost.
Clearly there is a concentration of flux in the teeth. The flux crossing the
airgap funnels into the teeth, which occupy approximately half of the
periphery at a radius taken half-way down the slots. This means that the
airgap flux-density will be of the order of one-half the maximum flux-
density in the teeth, i.e., at most, 0.8 T if the steel is kept below 1.6 T.
The flux-density in the magnets in a surface-magnet motor of the type
shown in Fig. 4.1 will be comparable to the airgap flux-density, or slightly
4-2
4. M a g n e t ic D e s ig n
higher. In motors of this type, only the so-called high-energy magnets
(Cobalt-Samarium; Neodymium-Iron-Boron) are capable of operating at
this high level of flux-density. Motors with ferrite magnets operate with
lower flux-densities, and of course they are less expensive,

4.2 Permanent magnets and magnetic circuits


A permanent magnet can be regarded as a flux source, and the magnetic
field can be calculated by means of the magnetic circuit, which is
analogous to a simple electric circuit with the following correspondence
between the variables:
Magnetic Circuit Parameter Electric Circuit Parameter
Flux [Wb] Current [A]
MMF [A-t] Voltage [V]
Reluctance [A-t/Wb] Resistance [Ohm]

T able 4.1 M a g n e t ic / e l e c t r ic c ir c u it a n a lo c y

Electric circuit analysis employs ideal current and voltage sources, and
real sources of current and voltage can be represented by their Thevenin
or Norton equivalent circuits, i.e. a voltage source in series with an
internal resistance, or a current source in parallel with an internal
conductance. (Conductance = 1/resistance). Similarly, in magnetic
circuits a permanent magnet can be represented by a Thevenin
equivalent circuit comprising an MMF source in series with an internal
reluctance; or by a Norton equivalent circuit comprising a flux source in
parallel with an internal permeance, Fig. 4.2. (Permeance =
1/reluctance). This internal permeance is sometimed called the magnet
leakage permeance.
The Thevenin and Norton equivalent circuits are exactly equivalent and
cannot be distinguished from each other by measurements at the
"terminals", since they both represent the same thing. Which one to use
is a matter of convenience. For example in the Norton equivalent circuit,
the internal leakage permeance represents flux that circulates inside the
magnet and does not emerge from the pole faces.
4-3
D e sig n o f bru sh less perm anent -m agnet m o t o r s

T H E V E N IN NORTON

MO

Fig. 4.2 Thevenin and Norton equivalent magnetic circuit]

The characteristics of a permanent magnet can be expressed graphically


in terms of the flux/MMF relationship at the "terminals" or pole-faces,
Fig. 4.3. This is analogous to the voltage/current relationship of an
electrical source at its terminals, and it is also quite similar to the
torque/speed characteristic of a brushless DC motor (Fig. 1.10). The
magnet can be "short-circuited by connecting a soft iron "keeper" across
its poles. This ensures that the MMF across the terminals is zero, and the
magnet is operating at the SHORT-CIRCUIT point in Fig. 4.3.

FLUX

Fig. 4.3 MMF/Flux characteristic of a permanent magnet of given dimensions

4-4
4. M a g n e t ic D e sig n
The "open-circuit" condition, on the other hand, requires that the flux
leaving the magnet poles is zero. In order to achieve this, an external
demagnetizing MMF must be applied to suppress the flux. The external
MMF must exactly balance the internal MMF of the magnet when the
flux emerging from the poles is zero. It is only possible to do this in a
magnetizing fixture with a separate DC coil which provides the external
MMF. The terminal MMF is negative because it opposes the internal MMF
Fc of the magnet, and is exactly equal to it. Fc is called the coercive MMF
because it is the MMF required to coerce the magnet to produce zero
flux. It directly expresses the resistance of the magnet to
demagnetization.
The amount of flux <I>r that can be produced into an infinitely
permeable keeper expresses the maximum available flux from the
magnet. >r is called the remanent flux. This is an historical term
describing how much flux "remains" in the magnet after it has been
magnetized. It should be interpreted carefully, because the ability of a
magnet to retain flux in a magnetic circuit depends on Fc as much as it
does on 4*r. It is better to think of 4>r as the flux "retained by a keeper
in the magnetic short-circuit condition.
In normal operation there is no keeper, and the magnet operates at a
flux below <&r- This is because the MMF drop across the airgap appears
as a negative, demagnetizing MMF as seen from the magnet "terminals".
In addition, the phase currents produce an additional demagnetizing
MMF which drives the operating point still further down the
characteristic.
It is clear from this that magnets require two parameters Fc and <t>r to
characterize them properly. Moreover, the slope of the magnet
characteristic relates Fc and For a given remanent flux <l>r, it is
desirable to have the flattest possible slope since this is associated with a
high value of Fc and a high resistance to demagnetization.
The most suitable magnets for brushless motors are the ferrites or ceramic
magnets, and the high-energy rare-earth and Neodymium-Iron-Boron
magnets. All these magnets have straight characteristics whose slope is
close to the theoretical maximum, and they are classified as hard magnets
because of their high resistance to demagnetization. Other magnets,
particularly Alnico magnets, have a high remanent flux but very low

4-5
D esig n o f brushless perm a nent -m a gnet m o t o r s

coercive MMF and low resistance to demagnetization. In the long history


of permanent magnet materials, spanning hundreds or even thousands
of years, it is only in the last two decades that truly hard permanent
magnet materials have been discovered and perfected. Twenty years ago
the "high coercivity" alloys referred to in the literature were far less
resistant to demagnetization than those available today.
The remanent flux $ f and the coercive MMF Fc depend not only on the
material properties but also on the dimensions of the magnet, and this
makes matters a little more complicated than the simple magnetic
equivalent circuit. The material property associated with $ r is the
remanent flux-density Br, and this is related to $>r by the equation
<*r = t 4 '1 )
where AM is the magnet pole area. The material property associated with
Fe is the coercive magnetizing force or coercivity Hc, and this is related to Fc
by the equation
= "cAv, (4-2)
where is the length of the magnet in the direction of magnetization.
From these equations it is clear that if the vertical axis of Fig. 4.3 is
scaled by 1/AM and the horizontal axis by there results a
relationship between By and i/M, the flux-density and magnetic field
strength in the magnet. These in turn are related to the magnet flux
and the MMF drop ^M at the operating point:
% = (4-3)
The graph of BM vs. HM is shown in Fig. 4.4, and this is, in fact, the
"second quadrant" of the B/H loop or hysteresis hop of the magnet
material, Chapter 16.
The magnets operating point generally moves reversibly up and down
the straight part of the characteristic in Fig. 4.4. This characteristic is
called the demagnetization characteristic. The slope is the recoil
permeability.
If B and H are in SI units ([T] and [A/m]), the magnetcharacteristics
are often plotted with ahorizontal axis \i^H instead of H.
4-6
4. M a g n e t ic D e s ig n

O P E N -C IR C U IT -
O P E R A T IN G P O IN T
NORMAL LOAD
DEMAG LOAD LINE
C H A R A C T E R IS T IC

/-P .C .
KN EE

H ca j Hc M ' ' m
PROJECTED
CURVE ACTUAL curve DEMAG EFFECT
OF PHASE CURRENT

Fig, 4.4 B /H characteristic of a hard permanent magnet material. This is the second
quadrant part of the full hysteresis loop. Note the krut. Also shown is the normal
operating point, the load line, and the effect of demagnetizing phase current.
This scales the units on the horizontal axis from A/m to Tesla, so that
the slope of the demagnetization characteristic becomes equal to the
relative recoil permeability, Hrec- Hard permanent magnets have a relative
recoil permeability in the range 1.01.1, close to that of air. With the
horizontal axis plotted as instead of H, a relative recoil permeability
of 1 has a slope of 45. When c.g.s. units (Gauss and Oersteds) are used
for B and H, there is no need to scale the horizontal axis because the
permeability of free space is unity in the c.g.s. system.
We have already seen that the airgap applies a static demagnetizing field
to the magnet, causing it to operate below its remanent flux-density. With
no current in the phase windings, the operating point is typically at the
point labelled OPEN-CIRCUIT in Fig. 4.4, with BM of the order of 0.7-0.95
* Bt. The line from the origin through the open-circuit operating point
is called the load line.

4-7
D e s ic n o f b r u sh l e s s p e r m a n en t -m a g n e t m o t o r s

S.I. c.g.s.
1 Tesla 10^ gauss or 10 kG
I A/m 471/1000 Oe
1 kj/m3 n/25 MGOe

T a b le 4.2 C onversion factors for macnetic units

The slope of the load line is the penneance coefficient, (P.C.). With the
horizontal axis plotted as instead of H, or in c.g.s. units, the
permeance coefficient is typically in the range 5-15.
When current flows in the phase windings, the additional field may drive
the operating point still further down the demagnetization characteristic,
depressing the airgap-flux-density as well as the magnet flux-density below
the open-circuit or no-load value. When the phase current is removed,
the operating point recovers to the open-circuit point, and the recovery
is complete and reversible provided that the excursion of the operating
point has not left the straight part of the demagnetization characteristic.
The straight part, over which the magnet normally operates, is called the
recoil line.
In Fig. 4.4 the intersection of the recoil line with the negadve-//M axis is
labelled //ca, the apparent coercivity. This is used later in the magnetic
circuit calculation. The actual coercivity is labelled Hc.
The best grades of hard permanent magnets have demagnetization curves
that remain straight throughout the second quadrant and in some cases
well into the third quadrant (negative BM as well as negative //M). These
magnets can withstand a demagnetizing field that is sufficient actually to
reverse the flux in the magnet, and still recover with no permanent loss
of magnetism.
Other materials have a knee in the second quadrant, as in Fig. 4.4. If the
operating point is forced below the knee, then when the demagnetizing
field is removed the magnet recovers along a lower recoil line. Fig. 4.4
shows an example in which the demagnetizing field is just sufficient to
reduce to zero. The magnet "recoils" along the depressed recoil line,
but it has lost almost 20% of its remanent flux. This loss is irreversible.
4-8
4. M a g n e t ic D esig n
A material which has a straight demagnetization characteristic at room
temperature may develop a knee in the second quadrant at higher
temperatures (this is characteristic of Cobalt-Samarium and Neodymium-
Iron-Boron magnets); or at lower temperatures (this is characteristic of
Ferrite magnets).
Generally, the remanent flux-density Bf decreases with temperature. This
effect is usually specified in terms of the reversible temperature coefficient of
Bt, quoted in % per degree C. If this coefficient is given the symbol a fir,
then the remanent flux-density at temperature 7"C is given by
B im = Brm M l + Br * I ? - 20)/100] (4.4)

where -B^o) 's value of Br at 20 C.


The cocrcivity also varies with temperature, but this is not as important
as the variation of the knee point. Unfortunately, published material data
rarely includes adequate information about the knee-point values //k and
ilj, and their variation with temperature. The design procedure followed
in this book begins by assuming a straight demagnetization characteristic,
equation (4.11), and at the end of the magnetic circuit calculation
checks that the magnet flux-density is greater than By by referring to
the actual magnetization curves of the material at the appropriate
temperature (Fig. 4.5). This procedure is safe and simple. The
temperature coefficient of H is not used, since it does not provide the
necessary information about the all-important knee point.
The effect of elevated temperature on the demagnetization characteristic
is usually to degrade the performance of the motor: firsdy by decreasing
the magnet flux and therefore the torque per ampere; and secondly by
requiring the operating current to be limited because of the upward
migration of the knee-point into the second quadrant. In most designs
the range of variation in the torque constant which can be tolerated is
quite small, of the order of a few percent. Limits are therefore set on the
motor temperature and on the phase currents to keep the operation
within this safe range. This explains why the very high temperature
properties of magnets (such as the Curie temperature or the temperature
at which metallurgical changes take place) are not usually of interest to
the motor designer.

4-9
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

Bm t
1.2
1.0
0.8

0.6

0.4

0.2

0
^ 0 HM T 1.2 1.0 0.8 0.6 0.4 0.2

Fig. 4.5 Magnet characteristics as z function of temperature

4.3 Approximate calculation of the flux


The first task is to identify the main flux paths and assign reluctances and
permeances to the various branches therein. Fig. 4.6 shows a simple
representation of the main flux paths in a four-pole motor. The main
flux or airgap flux <I>^ crosses the airgap and links the coils of the phase
windings. The magnet flux is the flux actually passing through the
magnet. The leakage flux 4>L is that part of the magnet flux which fails
to link the phase windings. These fluxes are all defined on a per-pole basis,
and their units are [WbJ. The leakage coefficient is defined as the
ratio of airgap flux to magnet flux:
/lkg = Z$ L = - I$* (4.5)
4>m + *L

4-10
4. M a g n e t ic D esig n

LEAKAGE FLUX

A IN FLUX

*9

Fig. 4.6 Main flux paths in surface-magnet interior-rotor motor


The leakage coefficient is less than 1, and its value depends on the
configuration of the motor. A typical "rule of thumb" value for most
motor types described in this book is 0.9.
Fig. 4.7 shows the magnetic equivalent circuit of one pole. Note that the
Norton equivalent circuit is used for the magnet. The leakage permeance

Fig. 4.7 Magnetic equivalent circuit of one pole

4-11
D esig n o f bru sh less perm anent -m a g n et m o t o r s
PL is in parallel with the magnet internal permeance The armature
MM F Fa due to phase current is shown as an MMF in series with the
airgap reluctance R^, but it will be assumed initially that = 0 (open-
circuit conditions).
In Fig. 4.7, permeances P and reluctances R are mixed freely, reflecting
the point of view in which a leakage permeance diverts magnet flux away
from the windings, while a reluctance (principally of the airgap) presents
a magnetic "resistance" to the flux. The reluctances of the steel stator
and rotor are omitted from Fig. 4.7 for simplicity, i.e., it is assumed that
the steel is infinitely permeable. The magnet permeance is given by

hi P

where PM is the pole-arc of the magnet in electrical radians, p is the


number of pole-pairs, is the magnet pole area, is the stack length,
and is the magnet length in the direction of magnetization. For
example PM = 0.9ti electrical radians is equal to 162 electrical degrees.
An electrical radian is p mechanical or ordinary radians.
The per-unit pole-arc or pole-arc/pole-pitch ratio is equal to Pm/ti and is
therefore 90% or 0.9 in this example. The radius rM is the effective
radius of the magnet, shown in Fig. 4.8 one-third of the way through the
magnet, measured from the inside radius of the magnet; however, a
safer approach is to use the actual inside radius of the magnet.
The magnet permeance PMf) is a highly idealised concept. It is derived
from the ratio r/F c and thus assumes a straight demagnetization
characteristic as well as uniform properties and magnetization
throughout the magnet.
The airgap reluctance R^ is given by

(4.7)

4-12
4. M a g n e t ic D esig n

M a g n e t (1 pole)

fig. 4.8 Dimensions for magnetic equivalent circuit calculation


where is the pole area of the airgap, taken at the radius rg midway
through the physical airgap. The airgap g' used in equation (4.7) is not
necessarily the physical airgap length g, but an effective value determined
by the use of Carters coefficient to allow for slotting. However, for
surface-magnet motors the airgap modification for slotting is small
because the magnet effectively acts as a large additional airgap, and
therefore it is acceptable to use g' = g.
It is evident from Fig. 4.7 that of the total remanent flux, only the
fraction Pg/^M o + + -Pg) crosses the airgap, with = 1/i^ ; thus
-'lkq (4.8)
LKQ^M0J
In terms of the respective flux-densities, since 4>r = and = 5gAg,
and making use of equations (1.5-1.7), wc get the following expression
4-13
D esig n o f brushless perm anent -m a g n et m o t o r s

for the open-circuit airgap flux-density Bg:


f am
LK0 A A
Bg - -----------------\ ~ fix a - fB ,' (4.9)

Note that having^ KG < 1 means that the airgap flux density is reduced
compared to the value it would have if there were no leakage. The
corresponding flux-density in the magnet is determined as

' B 7LKG
~ * 7 Ms r LKQV (410)

Since ^lkG < 1, for a given airgap flux the magnet flux and flux-density
are greater than they would be if there were no leakage. This is
intuitively correct, since the magnet must provide the leakage flux over
and above the airgap flux.
The operating point of the magnet can now be determined either
graphically, from Fig. 4.4; or by calculating //M from the equation which
describes the demagnetization characteristic:
Bh = + Bt ; > B^. (4.11)
The inequality B ^ > By, expresses the need to check that the operating
point is above the knee point.
With the foregoing equations it is not difficult to determine the value of
the permeance coefficient, and a convenient formula is
PC = - J - * h i * jli.. (4.12)
^UCG g'
In surface-magnet motors and the permeance coefficient is
roughly equal to / ^ / g . In order to achieve a high permeance
coefficient, desirable for operating as close as possible to the remanent
flux-density, the magnet length needs to be much greater than the airgap
length.
4-14
4. M a g n e t ic D e sig n
Another useful relationship involving the permeance coefficient [2] is

* * = tPC*
f T ;Mrec
,- x (413)
Since nrec is close to unity for most hard magnets used in brushless
motors, a high permeance coefficient ensures that the magnet operates
close to its remanent point. A value of 5 would be typical, giving ^ =
0.83flr with |ircc = 1. If the permeance coefficient is as low as 1, then
with prec = 1 5 ^ = Br/2, which corresponds to the maximum BH product
or energy product.
It can be seen from these design equations that, with a given magnet
material, the need for a high flux density is satisfied by making the
magnet as thick as possible in relation to the airgap length, while the
need for a large flux per pole is satisfied by increasing the magnet pole
area. For the open-circuit condition the magnet volume per pole can be
shown to satisfy the equation
2 W . ..
K, = -------I _, (4.14)
M I ^ mI
where is the magnetic energy per pole stored in the airgap, equal to
BgH^/2 x A x g. This energy is determined by the volume of the airgap
and the flux-density B , so in order to minimize the volume of magnet
material required, it appears that the magnet should be operated with
the maximum energy product If the demagnetization
characteristic is straight, then the maximum energy product occurs when
= Br/ 2, with a permeance coefficient approximately equal to 1 , i.e.,
the operating point is half-way down the demagnetization characteristic.
This theoretical result is never applied in practical motor design, however, because
of the allowances needed for the demagnetizing MMF of the phase currents and
temperature effects.
Nevertheless, it is still meaningful to talk about a magnet material as
having a high maximum energy product | b e c a u s e this is a single
number representing the fact that both the remanent flux-density and
the coercivity are high. In common parlance, the BIImax figure is widely
used to express the "strength" of various magnet grades, and the units
are usually MGOe (megaGauss-Oersteds) or kj/m .

4-15
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

4.4 Nonlinear calculation of the magnetic circuit


The magnetic circuit calculation in the previous section neglects the
MMF absorbed by the steel parts of the magnetic circuit, and is therefore
valid only when the flux-density levels are low. For motors using ferrite
magnets this calculation may be sufficient, but when high-energy magnets
are used it is generally necessary to take the steel into account and for
this a nonlinear calculation is necessary because of the saturation
characteristics of the steel (Chapter 16).
The nonlinear calculation described here uses Amperes Law to
accumulate MMF drops around the flux paths in Fig. 4.6. This includes
the outer loop of Fig. 4.6, which represents the main airgap flux, and
the inner loop which represents the leakage flux. The procedure is to
add up the MMF drops around the magnetic circuit, taking into account
the nonlinear B /H curve of the steel, and equate their sum to the
apparent open-circuit MMF of the magnet. This is defined as the product

^ca (4.15)
where Hca is the apparent coerchrity of the magnet. This is defined in Fig.
4.4, and is usually greater than the actual coercivity because of the knee
that may exist in the demagnetization curve in the second quadrant.
Like the linear calculation in the previous section, the nonlinear
calculation begins by assuming that the solution for will not fall below
the knee-point value 1^, and this must be checked independently.
Proceeding with the individual MMF drops, and starting with the airgap,
the airgap flux-density is initially assumed to be equal to the value
calculated in equation (4.9). Then
(4.16)

Assume that the stator yoke flux is equal to the gap flux crossing the
airgap over half the pole area. Then
(4.17)

4-16
4. M a g n e t ic D e sig n
where Asy is the yoke cross-section area, and

U sy - X s y (V
The functional notation represents a linear (or cubic-spline)
interpolation along the B/H curve of the steel. In other words, once the
yoke flux-density is calculated from equation (4.17), the magnetic field
strength H^ is determined from the B /H curve. Then

^ ^ * A* <419)
where L^. is the length of the flux-path through the stator yoke over
one-half of a pole-pitch. Similar equations are written for the stator teeth
and for the rotor yoke,giving MMF components Fst and respectively.
The magnet flux is taken to be
- -5 s- (4.20)
LKG

where fLKG is assumed known. Then

^ = MT 1 = M L K G (421)

and if equation (4.11) is used to find from we get

Fu *V (4-22>
Now all the MMF drops are added together

F - 4 + + F* + Fv> + (4-23)
The principle of the iteration is that if F > Fca, is decreased and the
calculation is repeated; if F< Fa , B is increased and the calculation is
repeated, and this continues until r is within 0.1% of F^. An under
relaxation factor can be used, multiplying the right-hand side of equation
(4.23) to assist convergence, or Newtons method can be used.
4-17
D esig n o f brushless perm a nent -m a gnet m o t o r s

4.5 Armature reaction and demagnetization


Current flowing in the stator tends to distort the magnetic field set up by
the permanent magnet. The larger the stator current, the larger the
distortion of the field. In DC motors this effect is known as armature
reaction. In brushless DC motors of the surface-magnet type, the effect
is weak because the magnets present a low permeability to the impressed
MMF of the stator current. In buried-magnet motors, however, the soft
iron pole shoes present a high permeability to the stator current and
consequently the field distortion is higher.
The most important effect of armature reaction in the brushless DC
motor is the possibility of partially or totally demagnetizing the magnets.
In normal operation the current is limited by the controller so that if the
magnet thickness and coercivity are sufficiently large, there should be no
risk of demagnetization. However, abnormal operating conditions can
produce large currents: for example, if full voltage is applied without
any current chopping when the rotor is stationary, the current is limited
only by the winding resistance. This current is the "locked-rotor" current
and is usually many times larger than the rated current.
An even worse case is possible if the transistors are switched 180 out of
phase with the correct firing angles, without current chopping, when the
rotor is rotating at maximum speed. In this case, maximum back-EMF
E is added to the supply voltage, and since these voltages are of the same
order, the resulting current (which is limited only by the winding
resistance) is of the order of twice the "locked-rotor" current.
Another operating condition that can give rise to large demagnetizing
currents is over-running, that is, when the DC supply voltage falls below
the back-EMF. This can happen when an over-running load drives the
motor faster than the no-load speed. The motor becomes an AC
permanent-magnet generator whose output is rectified by the
freewheeling diodes in the controller. As there is usually a large filter
capacitor connected across the DC supply, there is very little impedance
other than the winding resistance, and large currents may flow.
It is important to estimate the worst-case demagnetizing condition in the
magnet. The type of "manual" calculation described in this chapter gives
only an approximate estimate: in practice, worst-case demagnetizing
conditions are usually associated with marked saturation of the steel.
4-18
4. M a g n e t ic D esig n
Notwithstanding these reservations, a linear calculation is used and the
worst-case is assumed to be the locked-rotor condition with locked-rotor
current 7l r flowing. The stator ampere-tums ^ L R are assumed to be
concentrated between the inner and outer loops in Fig. 4.6. Then the
MMF-source appears in Fig. 4.7 as shown. Superposition is used to
calculate the flux and flux-density in the magnet:

= ^ mo ^*Ma ^ Ma. (4.24)


where is the normal open-circuit value of flux-density in the magnet,
given by equation (4.10), and is the additional component due to
the phase current. From Fig. 4.7,

M. (4 2 5 )
1 + LKG MO g

Using equation (4.6), this gives

x ^LKG __ ^ ^1rec^iojCLR 25)


An I + ^LKG^MO^g
The value of FLR depends on the locked-rotor current and on the layout
and connections of the winding. Fig. 4.9 shows the demagnetizing
conditions for a wye- and a delta-connected squarewave motor with the
rotor in two particular positions. In Fig. 4.9a, locked-rotor current ^LR
flows positively through phase 1 and negatively through phase 2, as is
normal for "two-phase-on" operation (Chapter 5). In the motor cross-
section the ampere-conductors of each phase appear as a 60 "phasebelt".
Because the axes of the phase windings are displaced by 120, the
adjacent phasebelts of phases 1 and 2 carry current in the same
direction, forming a combined phasebelt of 120 span, as illustrated by
the solid black segments labelled 1 and -2- On the opposite side of the
machine the return conductors form a 120 belt of current in the
opposite direction, shown hatched. A contour encircling the combined
120 belt of positive current is shown in Fig. 4.9a. Since the rotor and
stator steels are highly permeable, most of the MMF is dropped across
the two magnets and the two airgaps in series at A and B. This MMF is
*LR-
4-19
D esig n o f bhu shless perm a nent -m a cn et m o t o r s

Fig. 4.9 Conditions at locked-rotor, showing the demagnetizing ampcre-tums applied to


the magnet, (a) Wye connection (b) Delta connection
Fig. 4.10 shows how R is evaluated for a four-pole motor with one coil
per pole. Only the coils of phase 1 are illustrated. Each coil has Nc
turns, and the ampere-conductors in one phasebelt are those which are
enclosed within the contour, that is, 2jVc/ph/o, where / h is the phase
current at locked-rotor and a is the number of parallel paths through the
winding. In Fig. 4.106, a = 2. In Fig. 4.10c, a = 1.
4-20
4. M a g n etic D esig n

fig. 4.10 Electrical connections for calculation of winding MMF at locked-rotor


It is often convenient to express ^LR not in terms of Nc but in terms of
the total number of turns in series per phase N h, because this will
encompass the case where the number of coils/pole differs from 1
(including consequent-pole windings which have only 0.5 coils/pole: see
Chapter 5). If all the coils are in series then = Nc x Poles, but if
there are a parallel paths then Arph = Nc x Poles/a. In terms of the
number of pole-pairs p, then with two phases contributing to and
^LR(Y) = ^ph

= 2 phases * & * 2 Nc = 2 j"vph<?


0 A1 7
* * Ip v(4.27)/
_ ph_LRqo (2 airgaps).
P

4-21
D e sig n o f bru sh less perm a nent -m agnet m o t o r s

The locked-rotor current in the wye connection is approximately equal


to / lr^y) = where /?ph is the phase resistance.
The delta-connected motor is shown in Fig. 4.96. During one conduction
interval the DC current ^LR divides unequally between the two parallel
branches, with 2/3 going through phase 1 and 1/3 going through phases
2 and 3 in reverse. The ratio of the two branch currents is 2:1, i.e., the
inverse ratio of the branch resistances since one branch has a resistance
i?ph and the other has 2/?^. The resulting belt of ampere-conductors is
shown in Fig. 4.9b spanning 180. In the central 60 segment
corresponding to phase 1, there are Nph x 2/LR/3 ampere-conductors,
while in the outer two 60 segments corresponding to phases 2 and 3
there are A R/ 3 ampere-conductors. The total ampere-conductors in
the combined belt amount to (2/3 + 2 x 1/3) x A^h/LR or 4NPh W 3.
and in general with p pole-pairs
J P *W >- < 8>
The locked-rotor current in the delta connection is approximately equal
to 4 r(A) = 3 ^ / 2iiph.
The value of B ^ under load, determined from equation (4.24) with
equations (4.26-28), must not be lower than the knee value Bfc.
Consequently in computer-aided design programs it is common to
calculate the demagnetization current ^emag from these equations, as the
current that will just depress to the value B^. The controller must
then be set to limit the line current to Jjemag under all conditions. This
current is given by equation (4.27) for the wye connection and (4.28) for
the delta connection, and in both cases

^LR < ~ ( Mrec"o


~* +' ~ * <3*0 - 3c>. <429)

where ^ I0 is given by equation (4.13). This equation assumes that the


demagnetizing ampere-turns ,FLR are enclosed by an Amperes Law
contour that goes through two magnets (jf^j) and two airgaps (g) in
series.

4r22
4. M a c n e t ic D e sig n

q -axis

Fig. 4.11 Calculation of rotor leakage permeance


4.6 Calculation of rotor leakage permeance
The rotor leakage permeance PL can be calculated by simplified methods
of flux-path analysis or by finite-element analysis. A fuller discussion of
the significance of rotor leakage is given in Section 6.4.5. In this section
an example is studied, viz. the "spoke" type rotor of Fig. 6.1(c). Fig. 4.11a
shows a half-pole section of this motor, with three main fringing paths
egad, ADCS, and DEJC. Of these, ADCB and DEJC represent permeances
of rotor leakage flux which passes from the pole-piece back to the
adjacent pole-piece via the magnetic potential datum line Oq, the ^-axis.
On the other hand, egad represents the permeance of fringing flux which
crosses the airgap and increases the airgap permeance over the pole-face.
Permeance ADCSThis permeance can be calculated by assuming that the
flux-lines form concentric arcs, centre H. LetLAHB = 9. Consider an
4-23
D e sig n o f brushless perm a nent -m agnet m o t o r s

elemental strip bounded by two arcs, centre H, radii rand r+ dr, between
the equipotential surfaces AD and CD. The flux through this strip is
(4.30)
where /"is the MMF or magnetic potential drop between faces AD and
CD. The total flux between faces AD and CD is given by
(4.31)
The permeance ratio for this shape is the coefficient of (1/H qI^) fy/F, i.e.
(1/0) In (r2/ r j ) . In the case of the shape ADCB, rj = HA and r2 = HD.
Permeance DEJCThis permeance has an irregular shape that is not easy
to formulate analytically. The following technique is an amalgamation of
Roters technique [29] and the dual-energy method described by Prof.
Hammond [30]. Consider a brick-shaped volume carrying flux as shown
in Fig. 4.126. If the length of the brick (into the paper) is the
permeance is ^ nLs[k w/h, and w/h is the permeance coefficient. If this
is written as w h /$ , the product wh can be replaced by the transverse area
At and the permeance coefficient can be written At/ A2. The method
attributed to Roters involves applying this expression to any arbitrary
shape, such as the shape DEJC in Fig. 4.12a. The area At can be
calculated without difficulty if the shape is bounded by arcs and straight-
line segments, but it is not obvious what value should be given to /*a.
The crudest approach is to make h equal to the average of the two sides
that "channel the flux: in Fig. 4.12a, this would mean making k equal
to the average of the arc length DC and the straight-line segment EJ.
Other formulations of mean squares, or mean squares of reciprocals of
these sides, are possible, but if the shape is "reasonably square, or more
precisely, if it is close to a curvilinear square, then all these estimates give
more or less the same result.
In principle the estimated permeance coefficient can be improved by
means of the dual-energy method. In Fig. 4.126, the reluctance across the
area At (i.e., orthogonal to the original flux direction) is ( l A l 0Z.stk) h/w,
and the reluctance coefficient h/w can be written At /w~. The. same
uncertainty applies to the assignment of a suitable value to w2, but a
crude approximation is to use the average such as (DE+ arc CJ)/2. The
dual-energy principle states that the actual permeance coefficient

4
4. M a g n e t ic D e s ig n
estimate can be improved by replacing it with
4 (4.32)
A2 \
The closer the original shape is to a curvilinear square, the more
accurate this formula will be.
The two leakage permeance coefficients can be added to the permeance
coefficient for the segment of magnet that lies within the boundary of
the half-pole section, i.e. v>M/(L ^ /2 ). Equivalently, the per-unit rotor
leakage permeance can be written
Al = P\dcb + Pdejc + / ends (4.33)
whore s permeance coefficient for the shape ADCB, p ^ jC is
the permeance coefficient for the shape DEJC,and pendi is the per-unit
end-flux leakage permeance.
Permeance due to end-Jlux between pole-piecesFringing also occurs between
polc-pieces outside the active length. A very rough guide to the
contribution pcnAi is to assume that the flux flows in semicircular arcs
spanning the magnet, and use equation (4.31) with r-y = Lj^/2 and r2 =
2fj. Then
/ends * In . (4.34)
^ s tk rt
This formula includes both ends of the machine.
Permeance egad- 1This permeance can be calculated by equation (4.31) with
0 = ft/2, rj = he (airgap length), and r2 = hg. The resulting permeance
ratio can then be added to the permeance ratio for the half-pole section
of the airgap, which is anD /pg. It is common to augment the pole-arc
in this type of rotor by 2 x g / [gvD/Tj radians to account for fringing flux
that flows within the area hed at each edge of the pole-piece.
"Lumped" permeance calculations of this type have been used for a long
time and occasionally refined into complex nonlinear magnetic
ccjuivalent-circuit calculations. The method is less accurate and less
reliable than the modem finite-element method: its main virtues are
simplicity and ease of programming, and its use can be justified only

4-25
D e sic n o f bru sh less perm anent -m agnet m o t o r s

where extreme speed of calculation is required and accuracy is of


secondary importance, as (perhaps) in preliminary sizing calculations.
Some improvement in the method can be achieved by studying similar
cases with finite-element analysis to see the shape of the flux-paths and
to provide numerical checks on the accuracy, although this would
amount to abuse of the finite-element method in the eyes of many purists
in the field of numerical analysis! The need to be sure of the shape of
the flux-paths has always been the Achilles Heel of Roters' method, and
the dual-energy method does little more than attempt to make a good
estimate out of two poor ones.

4.7 Cogging
Cogging is the oscillatory torque caused by the tendency of the rotor to
line up with the stator in a particular direction where the permeance of
the magnetic circuit "seen" by the magnets is maximized. Cogging
torque exists even when there is no stator current. When visitors to a
trade exhibition rotate the shafts of brushless motors on the display
stands, they are feeling the cogging torque. When the motor is running,
additional oscillatory torque components can result from the interaction
of the magnet with space-harmonics of the winding layout and with
current harmonics in the drive current. These additional oscillatory
torque components are electromagnetic and are generally referred to as
torque ripple, while the term cogging is often reserved for the zero-current
condition. In a well designed motor the torque ripple and the cogging
should both be negligible, but it is possible for the torque ripple to
exceed the cogging torque by a large amount if the motor has an
inappropriate combination of winding layout, drive current, and internal
geometry. Manual rotation of a disconnected motor gives no indication
whatsoever about torque ripple.
One of the characteristics of a servomotor is low torque ripple, and "low"
generally means less than 1-2% of rated torque. This figure applies to the
combined effects of cogging and electromagnetic torque ripple, and is
met by the best quality sinewave-drive servomotors.
With a large number of slots/pole the cogging torque is inherently
reduced by the fact that the relative permeance variation seen by the
magnet is reduced as it successively covers and uncovers the slots one at
a time: indeed the permeance variation can be thought of as being
4-26
4. M a g n e t ic D e s ig n
concentrated at the edges of the magnet. A small amount of skew is then
usually sufficient to eliminate most of the cogging. When the number of
.dots/pole is closer to 1 , the slot geometry becomes more important, and
the widths of the teeth in particular can be adjusted to minimize the
cogging effect. An analysis of this approach was given by Ackermann et
al [1]. They developed an equation for the cogging torque:
^cog = j D L * E * K Jn ^ ( j g )
4 n n a L sk
where n = hS, k = 1,2,..., and S is the lowest common multiple of the
number of slots iVsloU and the number of poles 2p. The
s i n n o i stk^ function represents the effect of skewing, where o
is the skew angle, and An is the nth space-harmonic of the permeance
of the magnetic circuit "seen by the magnet as the rotor rotates.^, is the
nth space-harmonic of the magnet flux-distribution, and is the angle
of rotation of the rotor. This somewhat complex equation is derived from
the rate of change of coenergy as the magnet rotates. It clearly shows the
interaction between the space-harmonics of the magnetic circuit
permeance and the distribution of magnet flux, and can be used to
identify torque harmonics of a particular order.
The calculation of cogging torque from the rate of change of coenergy
can be applied with calculated values of coenergy obtained with the
fin ite -c le m e n t method. Because of the differentiation of the coenergy,
the finite-element solution needs to be very accurate, requiring a fine
mesh at the very least.
Cogging torque can also be compensated electromagnetically by adapting
the drive current waveforms to produce an electromagnetic torque ripple
component that cancels the cogging [2].
Other methods for reducing cogging include the use of birfurcated teeth
(as in Fig. 8.9), or punching holes in the tooth overhangs to modulate
the permeance variation [3 ]. Bifurcated teeth or "dummy slots" have a
similar effect to that of doubling the number of slots: the cogging torque
frequency is doubled and the amount of skew required to eliminate the
cogging is halved. Also, the permeance variation caused by uncovering
one half-tooth is of the order of half the variation caused by uncovering
a whole tooth, so the magnitude of the cogging torque decreases as well.
Tlie cogging torque can be adversely affected by partial demagnetization
4-27
D esig n o f bru sh less perm a n en t -m a g n et m o t o r s
of the magnets through over temperature or overcurrent, particularly in
surface-magnet motors because the degree of demagnetization is liable
to vary over the face of the magnet, producing a distortion in the back-
EMF waveform. This distortion tends to aggravate the torque ripple as
well as the cogging. Interior-magnet motors do not suffer from this
problem to the same extent
A summary of methods for reducing cogging torque is given in Table 4.3.
Increase airgap length
Use fractional slots/pole
Use larger number of slots/pole
Use thick tooth tipi to prevent saturation
Keep slot openings to a minimum
Use magnetic slot wcdges
Skew stator stack or magnets
Radius or chamfer magnet poles
Radius or chamfer stator tooth tips, or punch holes in tooth tips
Vary the magnetization of the magnet poles
Use bifurcated teeth
Use lower magnet flux-density
Compensate cogging by modulating drive current weaveform

T a b le 4.3 Methods f o r r e d u c in g c o g c in c t o r q u e

4.8 Retaining can losses


A cylindrical retaining can is often used to restrain the magnets of
interior-rotor, surface-magnet motors against centrifugal loading.
Alternatively glassfibre roving or other non-metallic wrap can be applied,
This technique permits very high speedsfor example, Prof. Takahashl
and his colleagues reported a 5 kW motor rated at 180,000 rev/min with
a fibre-reinforced plastic sleeve and a rotor surface speed of 200 m/s |4|.

4-28
4. M a g n e t ic D esig n

r.( . i .12
(BOD-
Magnet retaining can. (a) Geometry of induced current, (b) Short can haa
higher rcjistance to induced current, (c) Can divided into insulated rings
to reduce losses.

When a metallic can is used, eddy-cunrents are induced by variations in


the flux-density through the can thickness, and these cause additional
losses. The problem of calculating these losses was analyzed by Russell
and Norsworthy [5] in connection with screened-rotor induction motors.
Their approach is used here in simplified form to show the main trends
and effects of the main design parameters. Accurate calculation (and
measurement) of these losses is difficult especially in small motors.
Fig. 4.120 shows the geometry of the retaining can which rotates at N
rev/min and is penetrated by the magnetic flux-density B. The flux-
density is modulated by the slot openings as shown in Fig. 4.13.
4-29
D esig n o f brushless perm anent -m agnet m o t o r s

Fig. 4.13 (a) Dip in flux-density caused by stator slot opening as it moves relative
to the rotor at velocity v. (b) Construction of RMS value of the flux-
density variation in the can.

As the rotor passes a slot opening, the dip in the airgap flux-density
moves along the B-waveform which is otherwise moving synchronously
with the rotor. The variation in B is separated out in Fig. 4.136 and each
dip is represented as a half-sinewave of width 0 radians. The dips repeat
at intervals of A.s radians, where Xi = 2 n/jVs|otJ is the slot-pitch.
It is assumed that the eddy-currents are resistance-limited i.e., that they do
not modify the B-field. This assumption is appropriate bccause in cases
where retaining can losses are a problem, the eddy-currents need to be
resistance-limited in order to keep them low. (Bolton [6] presents a
method of analysis that does not make this assumption and allows the
eddy-currents to be resistance-limited or inductance-limited. His analysis
also includes a useful and relatively simple criterion for determining
when the eddy-currents are resistance-limited. This criterion is expressed
in terms of dimensions and can therefore be used as a design rule).
4-30
4. M a g n e t ic D e sig n
The rotation of the rotor causes an Afield in the can, given by v x B
where v is the surface speed in m /s. The corresponding current-density
in the axial direction is /= E/p. I f / 2p is integrated over the volume of
the can, the losses P can be calculated as well as the average loss per unit
area w. The result is
w = - * L ( B N P ft w / m 2 ( 4 .3 6 )
3600 p
The B value in equation (4.36) is the effective or RMS value of the flux-
density variation over the surface of the can. Referring to Fig. 4.136, this
can be estimated as
B = * iL (4.37)
V2 N
in IT). This result is intuitively derived as the RMS value of the half-
sinewave, modified by the effective "duty-cycle" implicit in the fact that
the dips repeat after rather than after p.
For example, if P/A^ = 1/4 and = 0.2 T, D = 100 mm, t = 0.5 mm, and
JV= 3000 rev/min, the surface loss is w = 3.6 W /cm 2 for copper, or 0.088
W/cm 2 for stainless steel (i.e., 0.57 W /in2).
The above analysis is "two-dimensional" and only considers current
flowing in the axial direction. Of course the current must flow in
complete loops as illustrated in Fig. 4.12. There is a circumferential
component as well as an axial component. Russell and Norsworthy
derived a simple formula for modifying the total can losses by a factor
to take account of the end-effects. Assuming that the can has the same
axial length as the rotor, i.e. LJ[k, the total can losses are given by
P = J^w A W (4.38)
where A = nDL^ is the surface area and
'PL*
Ks * 1 -
tanh
D (4.39)
P^&k
D

4-31
D esig n o f bru sh less perm a nent -m a g n et m o t o r s

Given that the stack length/pole-pitch ratio is 2pLlt^/nD , equation (4.59)


can be used to generate the examples in Table 4.4.
Stack Lcngth/pole-pitch
2 0.683
1 0.416
0.5 0.165

T able 4.4 F in it e -len g th effect fo r can lo sses [5]


For example, if a motor has "long" poles with Lstk equal to twice the
pole-pitch, = 0.683. If the can was split into four separate rings,
insulated from each other (Fig. 4.12c), the losses would be changed by
a factor 4 x 1 /4 x 0 .1 6 5 /0 .6 8 3 * 0.24, indicating a reduction of about
75% . The 1 /4 factor in this scaling arises from the fact that each ring has
only 1 /4 the original length in the axial direction.
The variation in the can flux-density represented by Bd in Fig. 4.13 can
be estimated by conformal transformation [4] or calculated by finite-
elements. Some examples of finite-element calculation are shown in the
airgap flux-density waveforms in Chapter 12.
The most effective means for reducing the can losses are indicated by
equation (4.36). The speed iVand diameter D are generally fixed, and
the minimum can thickness is determined by the strength requirement
This leaves B, the RMS variation in the flux-density through the can, and
p, the can resistivity, available to be varied in order to minimize the can
losses. The flux-density variation B can be decreased by increasing the
airgap or by reducing the slot opening. Reference [4] points out that
the slot opening is constrained to a minimum value by winding
requirements, and reports the effective use of an en la rg e d airgap to
reduce the effective B variation.
The can resistivity p for suitable materials such as Inconel, stainless
steel, etc., is generally high enough to keep the can losses acceptably low
at moderate speeds up to a few thousand rev/min, but for ultra-high
speeds (without direct cooling) it may be nccessary to use a non-metallic
material [4]. Magnetic steel may also be used provided it is thin en ou gh
to saturate and not short-circuit too much o f the magnet flux.
4-32
4. M a g n e t ic D e s ic n
In rare cases a motor may be fitted with a stationary can on the bore of
the statornot for retaining the magnets but as a septum to separate the
rotor space from the stator space in case there are toxic or dangerous
fluids flowing through one or the other. The stationary can is exposed
to the full magnet flux which alternates in polarity as the rotor rotates,
and its losses are liable to be at least an order of magnitude higher than
those in the rotor can.
The formulas given here for can losses are only indicative, and cannot be
expected to be very accurate. The can losses can be increased by PWM
harmonics in the stator MMF distribution.
Measurement of the can losses can be accomplished with a dynamometer
in which the motor is driven by another motor, with a very sensitive
measurement of the shaft torque. The SPEED Laboratory developed an
air-bearing dynamometer specially for this purpose. The test motor is
fitted with cans of various materials and thicknesses, and driven over a
range of speeds. The torque/speed curves are recorded. From these
curves is subtracted the "tare" of the torque/speed curve obtained with
no can. This "no-can" curve includes no-load core loss and windage and
friction loss. The resulting difference curve directly indicates the can
losses as a function of speed.

References
1 Ackemiann B et al [1992] Afeui technique for reducing cogging torque in a class of
bushiest DC. mot/m, IEE Proceedings 139, No. 4, 315-820.
2. Jouvc D and Bui D [ 1993] Torque ripple compensation inDSP-bavrihrushle.tt m o drive.
Intelligent Motion, PC1M Proceedings, Numberg 28-37
3 Gizaw D [1993] Permanent magnet bmxhtf.it DC. motor having reduced cogging, United
States Patent No. 5,250,867
4- Takahashi I, Koganezawa T, Su G and Oyama K [ 1993] A super high-speed PM motor
driven hy a quasi-current source inverter, IEEE Industry Applications Society Annual
Meeting, Toronto R57-W52
5- Russell RL and Norsworthy KH [ 1958] Eddy-currents and wall losvs in icreened-rotor
induction motors, IEE Proceedings 105A, 1B3-175
h Bolton H [19fi9] Transverse edge-effect in sheet-rolor induction motors, IEE Proceedings
116, No. 5, 725-731
4-33
I
5. ELECTRICAL DESIGN
j .l Introduction
Now that we have reviewed the configuration of the motor and
determined how to calculate the magnetic flux, we are in a good position
to study the basic electrical operation. This chapter considers the
windings in more detail: in particular, the factors affecting the layout of
the windings and the calculation of the correct number of turns. This
requires a simple explanation of the operation of the ideal brushless DC
motor, describing how a squarewave EMF is generated and how this
relates to the operation of the power electronic controller. Motors with
1,2, and 3 phases are considered, and there is a discussion of the use of
wye and delta connections in three-phase motors. Although this chapter
is mainly about squarewave motors, it deals with many fundamental
points which are important for sinewave motors also (see Chapter 6).
The chapter includes a treatment of winding inductances and the analysis
of slodess windings.

5.2 Basic windings


5.2.1 Squarewave motor
A simple concept machine' is shown in Fig. 5.1. The 2-pole magnet has
a pole arc of PM = 180 and the airgap flux-density distribution is ideally
a square wave as shown in Fig. 5.2a. The magnet flux crosses the airgap
radially, and Fig. 5.2a is interpreted as the variation in the radially-
directed flux-density around the circumference of the airgap. The
angular coordinate 6 is used to measure angle around the rotor from the
rf-axis. Thus the magnet flux distribution in the airgap is represented by
the function 5(0). In practice, fringing causes the comers to be
somewhat rounded (see Chapter 8). The filled magnet is the S pole and
the unfilled one is the N pole. The flux from the N pole is positive (i.e.,
directed radially outward, while the flux from the S pole is negative
(radially inward).
The rotor position is defined with reference to the rotor reference axis
or direct axis (d-axis), which is taken as the centre of a S pole. The rotor
position is the displacement of the ttaxis from the reference axis (the
5-1
D e sig n o f bru sh less perm anent -m agnet m o t o r s

*-axis). In Fig. 5.1 the rotor is shown at the position \ = 0: i.e., the d-axis
coincides with the x-axis. The *-axis is the stator reference axis and for
this reason the slot-numbering scheme begins with the axis of slot 0 on
this axis. Slot 0, of course, is the same as slot 12 in this machine.

d-AXIS

Fig. 5.1 12-slot, 2-pole brushless DC motor showing the winding of phase 1 with 1
coil/pole, 5/6-pitch, and 2 slots/pole/phase
When the rotor moves, the whole flux-distribution moves. Thus, for
example, Fig. 5-26 shows the distribution B(9) when the rotor is at the
position =105.

5-2
5. E le c t r ic a l D esig n
The stator has 12 slots and a three-phase wye-conected winding. Only
phase 1 is shown. There are 6 slots per pole and 2 slots per pole per
phase. This is not necessarily a desirable combination in practice (see
Chapter 3) but it is useful for illustrating the winding principles
developed in this chapter for both squarewave and sinewave motors.
Each phase winding consists of two coils 1 and 2 of jVj turns each. Coil
1 is wound in slots 1&6 and thus has a span of 5 slot-pitches. The span
is also called the pitch or throw of the coil. Since one pole-pitch is 6 slot-
pitches in this machine, the coil pitch is said to be 5/6. The "start" of
coil 1 is in slot 1 , and its axis is located at half the span further round
the stator, i.e. at 5/2 = 2-5 slot-pitches from the start. Since slot 1 is
located at 30 (one slot-pitch) from the *-axis, the axis of coil 1 is at 30
+ (5/2) x 30 = 105 from the #-axis.
The flux-linkage of a stationary coil is represen ted by th e function i|>c (),
because it varies as the rotor rotates. Fig. 5.2c shows the waveform of
the flux-linkage i|/cl() and Fig. 5.2d shows the back-EMF ci() in coil
1 as the rotor rotates. These waveforms are plotted vs. rotor position
The origin is at 5 = 0t corresponding to the particular rotor position
shown in Fig. 5.1. Thus the negative peak of the flux-linkage i|rc] occurs
when the rotor is at the position = 105, where the S-pole rf-axis is
aligned with the axis of phase 1.
The back-EMF is derived directly from the flux-linkage waveform by
Faradays Law:
* 5) . (5.1)
df oc
The notation e() simply means the EMF plotted as a function of rotor
position rather than plotting it as a time waveform e{t). The waveshape
is unaffected because E, = and Ci>ra = d^/dt. This equation applies
not only to individual coils but also to the complete phase winding.
The corresponding waveforms for coil 2 are identical but opposite in
sign, because coil 2 is displaced by 180 and therefore links magnet flux
of exactly the same magnitude but of opposite polarity to that linked by
coil 1. When the coils are connected in series, coil 2 must be effectively
reverse-connected so that the total back-EMF is doubled to 2ecl, otherwise
the total EMF would be zero.

5-3
D esig n o f bru sh less perm anent -m a c n et m o t o r s

6(5=0)
a

B (? = 105)

CONDUCTION INTERVAL

ec2<5>
e
Wi 2//( /lM) W ///A
ec3< S>

's(5>

180 360

Fig. 5.2 (a,b) Magnet flux distribution (c) flux-linkage waveform and (d,e,f) EMF and
current waveform* corresponding to the motor of Fig. 5.1.
5-4
5. E l e c t r ic a l D e sig n
In Fig. 5.1 the conductors of phase 1 occupy 4 of the 12 slots, leaving 4
slots for each of the other two phases without requiring any slot to
contain conductors from more than one phase. A single-layer winding is
therefore built up with phase 2 in slots 5&10, 4&11; and phase 3 in 3&8,
2&9. This ensures that the winding axis of phase 2 is 120 ahead of the
phase 1 axis (i.e., at 105 + 120 = 225) and the winding axis of phase 3
is 240 ahead at 105 + 2 x 120 = 345. The phase 2 axis is along the
centre-line of tooth 7/8 and the phase 3 axis is along the centre-line of
tooth 11/12. Thus the phase winding axes are displaced from each other
by 4 slot-pitches ( 120).
With a magnet pole-arc of PM = 180 and a coil-pitch of 5/6, the back-
EMF is ideally flat-topped over 5/6 of a pole-pitch or 150. However,
fringing effects at the edges of the magnet poles cause the EMF
waveform to be rounded and the effective flat-top may extend over no
more than 120 in practice. This is 1/3 of a revolution. Now phase 1 can
produce constant torque if it is fed with constant current of the correct
polarity during the 120 interval. If phases 2 and 3 are fed with the same
current during the other two 120 intervals, constant torque can be
produced through one complete revolution, as shown in Fig. 1.6.
The necessary phase current waveforms are shown as blocks in Fig.
5.2d,e,f It is important to note that the conduction intervals are 120
wide, and alternate in polarity. The resulting current waveform in each
phase is described as a "120 squarewave1', and the three phase currents
form a balanced set with equal phase displacements of 120. Becausc of
the phase displacement between phases, there is a commutation
(transistor switching) every 60, and therefore there are 6 commutations
per cycle of the fundamental frequency. There are two phases
conducting at any and every instant, and this is often called 2'phaseon
operation. A commutation always switches one phase off at the same time
as another one is switched on. The sequence of switching the six power
transistors in Fig. 1.8 is shown in Fig. 10.3 together with the three phase
current waveforms for a wye-connected stator. For a delta-connected
stator, the phase and line current waveforms are shown in Fig. 10.4; the
transistor switching sequence is identical for wye and delta conections,
and in fact the controller does not "know" the difference between them.
Note that in Figs. 10.3 and 10.4, the origin is taken as the axis of phase
1, i.e. at the 105 point in Fig. 5.2.

5-5
D esig n o f brushless perm a nent -m a g n et m o t o r s

PH A SE 1 A XIS

d-AXIS

Fig. 5.3 12'ilot, 2-pole brujhleu DC motor showing the winding of phase I with 2
coils/pole, having pitches of 5/fi and 1/2 respectively.
5.2.2 Effect of additional coils
Fig. 5.3 shows the phase 1 winding with two additional coils 3 and 4, each
of which has a pitch of 3 slots. This is 1/2 of a pole-pitch, or 90 electrical
degrees. The axes of coils 3 and 4 are coincident with those of coils 1
and 2 respectively, so their flux-linkages and EMFs will be in phase with
i|cl and ec]. The additional coils are assumed to have the same numbers
of turns as coils 1 and 2. Fig. 5.4 shows the flux-linkage i|jc3 and EMF
of coil 3 together with t|rcl and ecl, as well as the total flux-linkage and
5-6
5. E l e c t r ic a l D e s ig n

COIL e 150
EMF c3

e
FUNDAMENTAL
ACTUAL 90

e c 1 + e c3
/
/ s.
s
\
s

J ^ __

30 deg

180 360

Eft- 5.4 (a,b,c) flux-linkage waveforms and (d,e,f) EMF waveforms corresponding to the
motor of Fig. 5.3.
5-7
d e s ig n o f bru sh less perm a nent -m a gnet m o t o r s

FUNDAMENTAL

Fig. 5.4g Expanded view of Fig. 5.4/ showing the actual EMF waveform and the
fundamental harmonic from Fig. 5.4/

EMF obtained by connecting coils 1&3 in series. Connecting coils 2&4 in


series with coils 1&3 merely doubles the resulting EMF, provided that the
polarities are correct.
With a magnet pole-arc PM of 180, the maximum flux linked by any coil
is proportional to its span. Consequently the maximum flux linked by
coil 1 is only 5/6 of that which would be linked by a full-pitch coil. A full-
pitch coil would have a perfecdy triangular flux-linkage waveform with no
flat top, and a perfecdy square EMF waveform with a flat top 180 wide.
Coil 3 links a maximum flux only 1/2 that of the full-pitch coil, i.e. 3/5
of the maximum flux linked by coil 1. If coils 1 and 3 have the same
number of turns, then their flux-/mAagej are in this ratio. This is
indicated by the dotted lines in Fig. 5.4a. The peak back-EMF in coil 3
is the same as in coil 1 , because although the change of flux-linkage is
only 3/5 that of coil 1, the change takes place over 3/5 of the angle.
When the effect of fringing is included, the comers in both the flux-
linkage and EMF waveforms become rounded. The effect on the EMF ecj
+ ecS is shown in Fig. 5.4g-by the dotted line labelled ACTUAL, and this is
the so-called "trapezoidal" back-EMF waveform. One of the effects of
fringing is to make the flat top narrower. In Fig. 5.4 it is clearly less than
the 120 required to maintain constant torque, and this waveform is
apparently less than ideal for a wye-connected squarewave motor,

5-8
5. E l e c t r ic a l D esig n
especially if smooth torque is important.
On the other hand, adding the additional coils 3 and 4 increases the
number of steps in the overall back-EMF waveform, and although it may
not be strikingly obvious from Figs. 5.2/and 5.4/ the waveform of the 4-
coil winding has a smaller harmonic content and is closer to a sinewave.
The fringing (and also any skew) further reduces the harmonics, and it
can be correctly inferred that for a sinewave brushless AC motor the
additional coils are beneficial, because the production of smooth torque
requires the combination of sinewave back-EMF with sinewave current
(Chapter 6).
The back-EMF waveform ecl appears to lead the corresponding flux-
linkage waveform i(rcl by 90: this is also true of the fundamental
harmonic components T t and Ecl, which are phasors and are therefore
related by the equation
^1 = <52>
where o) = 2ti/ and / is the supply frequency. This equation does not
apply to the actual waveforms, because they are not pure sinewaves: it
applies only to the fundamental harmonic components.
5.2.3 Lap windings and concentric windings
The manner in which coils 1-4 are wound, inserted, and connected
together is extremely important from a manufacturing point of view,
where the cost of winding is important. The addition of coils 3 and 4 in
the previous section causes phase 1 to occupy 8 of the 12 slots, so that
when phases 2 and 3 are added every slot contains two coil-sides: i.e., a
double-layer winding. This is more difficult to wind than the single-layer
winding of Fig. 5.1.
Fig. 5.5 shows the principle of a lap winding corresponding to Fig. 5.3. In
the lap winding, all roils are identical in pitch and number of turns.
When the coil pitch is greater than 1 slot pitch, the concentric winding is
common. It can be seen from Fig. 5.5 that the flux-linkages and back-
EMFs of both windings are identical because the disposition of
conductors is identical, and so is their electrical connection (i.e., all in
series). The windings differ only in the end-region, by having slightly
different resistances and inductances in the end-windings.
5-9
D esig n o f bru sh less perm a nent -m agnet m o t o r s

WINDINGS

3 ACTIVE
LENGTH

(a) LAP

Fig. 5.5 Lap arid concentric windings.

Big. 5.6 Winding diagram for motor of Fig. 5.1

5-10
5. E l e c t r ic a l D esig n
A "developed" winding diagram for the motor in Fig. 5-1 is shown in Fig.
5.6, for phase 1 only. The other phases are similar. For coil-winding, the
route taken by the wire is designed to minimize the length and number
of interconnects in the end-windings. This helps to minimize the volume
of copper in the end-windings and to minimize the winding resistance.
$.2.4 Multiplepole machines
So far the discussion has been confined to a 2-pole motor, but the same
principles apply to multiple-pole motors. The waveforms in Figs. 5.2 and
5.4 remain unchanged if 0 and are in electrical degrees. TTie 4-pole
equivalent of the motor in Fig. 5.3 has 24 slots, so that the number of
slots/pole is the same. The coils are wound with the same pitches as a
fraction of the pole-pitch, which is reduced to 90 actual degrees (= 180
electrical degrees). The period of the waveforms in Figs. 5.2 and 5.4
covers 180 actual degrees {= 360 electrical degrees). An angle in electrical
degrees is p times the same angle in actual degrees, (p = poU-pairs).

Kg, 5.7 Consequen t-pole arrangement of phase winding and/or magnets.

5-11
D esig n o f bru sh less perm a nent -m a gnet m o t o r s

In the 2-pole motor of Fig. 5.3 there is one coil-group per pole, so that
with 2 poles there are 2 coil-groups and 4 coils in each phase. If the 24-
slot motor retains 1 coil-group/pole there will be 4 coil-groups and 8
coils per phase. The additional 2 pairs of coil-groups can be connected
in series with the first group; or in parallel, giving 2 parallel paths through
the phase winding. This multiplies the back-EMF at the phase terminals
by 50%, and doubles the phase current required to produce the same
torque. In other words, for a fixed number of coils the back-EMF
constant ftg and the torque constant kj (torque/amp) are both inversely
proportional to the number of parallel paths.
5.2.5 Consequent-pole vnndings and magnets
In the 4-pole motor it is not mandatory to have 4 coil groups per phase.
An alternative arrangement is to have only two coil groups diametrically
opposite each other. This is called a consequent-pole winding. Such a
winding is shown in Fig. 5.7 with a 12-slot stator. The 4-pole flux pattern
could be produced by any one of the four configurations acting alone:

The 4-pole magnet, as shown in Fig. 5.7 Non Consequent-pole


A magnet with any 2 opposite poles removed, Consequent-pole Magnet
e.g. N-O-N-O or S-O-S-O
The phase winding as shown in Fig. 5.6 Non Consequent-pole
A phase winding with additional coils in slots Consequent-pole winding
5&S, 11&2 ------------------ ------------------------- 1
T a b l e 5.1 C o n s e q u e n t -p o l e w in d in g s

Either or both the windings and the magnet may be of the consequent-
pole configuration. The consequent-pole configuration is simpler to
assemble and can save copper or magnet material. For example,
removing the two N magnets would leave the flux pattern essentially
unchanged, but the airgap flux-density and the flux per pole would be
slighdy reduced because each magnet is now forcing flux through two
airgaps instead of only one. The thickness of the magnet may need to be
increased to compensate for this. The increase can be worked out from
equations (4.9) and (4.10). In order to restore the remaining magnets

5-12
5. E l e c t r ic a l D e sig n
to the same operating point, their thickness would have to be
doubled. However, it may be that when all four magnets are used the
thickness is greater than it needs to be, owing to the fact that there is a
minimum thickness to which magnets can be manufactured. If this is the
case, then the consequent-pole arrangement makes better utilization of
the magnets even though they may be working at a slightly lower
permeance coefficient and flux-density.
5.2.6 Computer-aided design of windings
The PC-BDC computer program includes a winding editor which is shown
in Fig- 5.8 with the same winding as in Fig. 5.7. The winding editor can
be used to construct any combination of coils, or alternatively it can
construct many "standard" types of winding automatically. Whatever the
distribution of coils, the program can subsequently calculate the EMF
waveform as well as the resistance and inductance of each phase and the
mutual inductance between phases. This is especially useful when the
winding is non-standard or very complicated. A winding-construction
algorithm is described in Section 3.8.3.
PC-BDC 3 .1 Cross-sect Ion editor CC) 1992 TJOI.Wfce

Locate Go Coil side


Ho TurnsGo Ret Spai

Z1 12 Z
12 0
5
11
3
3

CLU, All phases


SlotHo - 1
Save flu It Go Help

Fig. 5.8 Screcn-capture of

5-13
D esig n o f bru sh less perm a nent -m a g n et m o t o r s

In squarewave motors it is important to ensure that the EMF waveform


has a sufficiently wide flat top. For simple configurations this can be
checked by hand, but if the designer is to have complete freedom to
explore combinations of slot number, winding pitch, number of coils,
magnet arc etc., then a computer analysis is virtually essential.

5.3 Wye and delta connections


5.3.1 Wye connection, with 12CF trapezoidal phase EMF waveform
The circuit diagram for wye connection of three phase windings is shown
in Fig. 10.1. The labelling of the transistors is such that for squarewave
operation they commutate in the sequence 1,2,3,4,5,6,1... and there are
two transistors conducting at any one time. For sinewave operation there
are generally three transistors conducting at any one time.
The lines are labelled A,B,C and the phases are labelled 1,2,3. In the wye
connection the line currents are identical to the phase currents, i.e.
= *1> *B = h an<* t = % The line-line voltages are = t/j - u2, t^s = Uj
- vs, and vsi = u3 - Vj. The double-subscript notation for voltages is used
rigorously in this fashion.
A complete set of line currents, line-line voltages, and phase voltages is
shown in Fig. 5.9a for the idealised squarewave wye-conected motor
which has phase EMFs and that are 120 trapezoids. These
waveforms would be produced by the stator of Fig. 5.1 with a magnet
having a pole-arc of 180. The ideal phase currents il, and ij are 120
squarewaves in phase with the respective EMFs, as shown by the
rectangular blocks in the upper traces in Fig. 5.9a. This means that the
line currents are also 120 squarewaves. The waveforms for phases 2 and
3 are grouped to show their symmetry with respect to phase 1. The phase
1 waveforms are singled out as representative of all three phases.

Since the EMF is synchronous with the rotor position, the shaft position
sensor ensures the synchronism between the EMFs and the line currents,
even though there is usually a voltage drop in the winding resistance and
leakage inductance. This voltage drop means that under loaded
conditions the terminalvoltages are notin phase with the back-EMFs, and
do not have quite the same waveform.

5-14
5. E l e c t r ic a l D esig n

0 30 60 90 120 150 180 210 240 270 300 330 360


e i sm m r

JSl'.vi.v.v?

1
\

OJ
m m i - - 1*......... l / 3

e12

23 \ / e31
V

\ /
:V :.
/
N X y
/
/

fig. 5.9a Idealised squarewave motor : wye connection, 120 flat-top EMF
Waveforms

The 120 squarewave currents require a commutation every 60, and the
switching sequence that achieves this is given at the bottom of Fig. 5.9a.
5-15
D e sic n o f bru sh less perm a nent -m agnet m o t o r s

0 30 60 90 120 150 180 210 240 270 300 330 360

e12 x 112
j j N y^ n '31M s
K

t
\

// m
\\

Q5 Q1 Q3 Q5

Q6 Q2 Q4 Q6

Fig. 5.9b Idealised squarewave motor : wye connection, 50* flat-top EMF waveforms
5. E l e c t r ic a l D e s ig n
In any 60 interval, the basic electromagnetic torque production is given
simply by

where eis the "reigning" line-line voltage (i.e., the one connected to the
supply through the two conducting power transistors), and i is the line
current. Since there is only one conduction path through the winding,
there is no ambiguity about the meaning of e and i For example, during
the period 30-90 e = and * * ip with transistors Q1 and Q 6
conducting. If all the phase waveforms are assumed to have a peak value
of 1 unit, then the peak line-line EMF is 2 units and the peak line
current is 1 unit, and if wm is taken to be 1 unit then the torque during
this 60 interval is 2 x 1 /I = 2 units.
The current ij during the 30-90 interval flows in the positive direction
in phase 1 and in the negative direction in phase 2, with transistors Q 1
and Q 6 conducting. This current can be regarded as a loop current and
it is expressed by the notation and similarly for i >3 and ij, in later
intervals (see Fig. 5.11). During every 60 interval the same conditions
prevail, the only difference being that is substituted in turn by eJ2,
while i is substituted in turn by ij2> *23, _*i2>The alternation
(i.e., polarity reversal) of the line-line EMFs is natural, being caused by
the passage of the N and S magnet poles. The alternation of the loop
currents is forced by the commutation process. Note that during the
commutation from i,2 to -ijj, remains unchanged but ig switches off
and -ij switches on.
A "block" of the torque waveform is shown immediately below the line
current waveform i 12 in Fig. 5.9a, labelled e12 x ij2- Its amplitude is 2
units and it remains constant throughout the 60 conduction interval
because e12 and i12 are both constant during this interval, and there is
no other current in the machine.
During the next and subsequent intervals the same block of torque is
replicated but with a different set of EMFs and currents. From 90-150
it is (~<3j) x (>3i)> then from 150-210 it is e2g x *23>etc- Consequently
ihe electromagnetic torque remains constant from each conduction
interval to the next, as shown in the lowest trace in Fig. 5.9a.

5-17
D esig n o f bru sh less perm a nent -m a g n et m o t o r s

The torque can also be regarded as being produced simultaneously by by


the individual phases rather than by the succession of single conducting
loops. The waveform labelled Tj in Fig. 5.9a is the product of the phase
EMF and the phase current ij, and similarly for T2 and T3 from
phases 2 and 3. The resulting torque blocks are 1 unitin magnitude and
120 wide, corresponding to the fact that each individual transistor
conducts for 120, and so does each phase. The total torque is the sum
T= Tj + Tg + and this has the constant value of 2 units since only two
phases are producing torque at any instant.
In the first half of the phase 1 conduction period, phase 2 produces an
equal constant torque block of 1 unit magnitude, while phase 3 is off.
During the second half of the phase 1 conduction period, phase 2 is
replaced by phase 3, i.e. -i, is replaced by while ij continues
unchanged.
This set of waveforms is extremely important because it shows that in its
ideal form the brushless DC squarewave motor can produce absolutely
constant torque when constant current is drawn from the supply. This
property is not shared by all machines. It is normally associated with the
classical machinesthe DC commutator motor (with pure DC), and the
polyphase AC synchronous and induction motors (with pure AC). The
torque is, of course, proportional to the current, which makes the
machine particularly well suited for control applications and amenable
to linear systems control theory (Chapter 14).
In Fig, 5.9a it appears that there is a 30 phase shift between the
fundamental of the line-line EMF and the fundamental of the phase EMF
(e.g., between e1 and e12)- This is the same as in the sinewave motor,
where ideally the fundamental is the only non-zero component, and
indeed this is a familiar characteristic of three-phase AC circuits.
Fig. 5.13a shows the operation of this motor in a very compact manner.
The motor is shown with 2 poles but the principle applies with any
number of poles. The stator winding is represented by phasebelts of
ampere-conductors. With three phases, and two poles, the distribution of
stator ampere-conductors is shown with the direction of the current
indicated by the appropriate shading. For wye connection the two
conducting phasebelts combine to produce a continuous 120 belt of
positive ampere-conductors, and a continuous 120 belt of negative

5-18
5. E le c t r ic a l D e s ig n
ampere-conductors, leaving two unexcited belts each 60 wide. A
winding which will produce this pattern is the one shown in Fig. 5.1 and
discussed in secdon 5.1.1. Note that this winding is not truly
concentrated, became the conductors of like polarity are "distributed"
over a 60 arc.
The rotor is shown as having a 180 magnet pole-arc, and it is tacitly
assumed that the magnet is radially magnetized, to produce a perfectly
rectangular flux distribution around the airgap. With a 60s phasebelt,
this magnet pole-arc produces phase EMFs with a 120 flat-top.
The diagram in Fig. 5.13a shows the principle of the squarewave motor
in a particularly graphical way. It can be assumed that positive
contributions to the torque are produced at all angles where there is
overlap between magnet and ampere-conductors of like polarity. There
is an arc of 120 of overlapping N-pole with one ampere-conductor
polarity, and 120 of S-pole with the other ampere-conductor polarity.
The position shown in Fig. 5.13a corresponds to 60 in Fig. 5.9a. At this
position the rotor can rotate 30 forwards or backwards with no change
in these overlap angles. At the end of this interval (90) there is a
commutation, which "resets" the overlap pattern, whereupon the next
"conduction interval" begins. The constancy of the torque follows from
the fact that the magnet pole-arc exceeds the ampere-conductor belt
width by 60.
Fig. 5.13a also shows that the ideal phase EMF has a 120 flat top, as this
is the difference between the magnet pole-arc and the phasebelt arc.
This is perhaps not immediately obvious, but it becomes clear if the BLv
waveforms are plotted and added together for all the conductors in the
phasebelt, especially if they are considered to be uniformly distributed
through the phasebelt.
The effect of reducing the width of the flat top of the phase EMF
waveform can be seen in Fig. 5.96 which shows the effect of a 60 flat-
topped phase EMF waveform. This waveform would be produced with
the winding of Fig. 5.1 if the magnet pole-arc were reduced from 180
to 120. The line-line EMF now has no flat top at all, and consequently
the torque production is not constant throughout the 60 conduction
interval. In fact, the peak-peak torque ripple is 28.6% of the mean
torque.
5-19
D esig n o f brushless perm anent -m agnet m o t o r s

In practice it may be desirable to use a 120 magnet arc for


manufacturing or cost reasons. To eliminate the torque ripple while
retaining a wye-connected winding, the coil pitch could be reduced to
2/3 instead of the 5/6 shown in Fig. 5.1. To understand why this works
it helps to think in terms of the relative overlap angles in Fig. 5.13. The
width Tph of the flat top on the phase EMF waveform in electrical
degrees or radians is given by
r ph = I 0 - M (5-4)
where by is the phasebelt angle. In Figs. 5.1 and 5.13a, by = 60 and (3^
= 180, so T , = 120. When two phases are conducting in the wye
connection, the effective ampere-conductor belt is increased to the
"combined" value because the phase belts of the two phases are
adjacent or overlapping. In Figs. 5.1 and 5.13a, 6^ = 26j = 120 (4 slot-
pitches) because the conductors of phase 1 in slots 12 and 1 are adjacent
to the conductors of phase 2 in slots 2 and 3, without overlapping. The
line-line EMF is the difference between the phase EMFs, and the width
IY l of its flat top is consequently
I'll = <5-5)
With the winding shown in Fig. 5.1 and pM = 180, = 60, as in Fig.
5.9a. If the magnet pole-arc is reduced to 120, then T h = 60 and
= 0, as in Fig. 5.9b, producing maximum torque ripple.
If the coil span is reduced to 2/3, the effective value of by increases to
120. The effective value of b^ is 1.5 x by because there is now 50%
overlap between adjacent phasebelts. If the magnet pole-arc is reduced
to 120 then with b^ = 180 the ampere-conductor distribution has an
effective belt width that exceeds the magnet pole-arc by 60. This is
shown in Fig. 5.136, which shows that the conditions for smooth torque
are restored.
When the coil span is reduced to 2/3 in the motor of Fig. 5.1, the three-
phase winding occupies only every other slot, and the slot number could
be reduced to 6.
Another way to achieve smooth torque with a 120 magnet and a wye
connected stator winding is to add coils as in Fig. 5.3. The phasebelt
5-20
5. E le c t r ic a l D esig n
angle 6j increases from 2 slot-pitches (60) to 4 slot-pitches (120). At
(he same time the combined phasebelt angle becomes 1.5 x 120 =
180 (6 slot-pitches), as shown by the shading in Fig. 5.136. When phases
1 and 2 are conducting, every slot contains two coilsides and a current
of 2NT, where Nc is the number of turns per coil. The ampere
conductors in slots 11,12,1-4 are all in the same direction, and the
ampere-conductors in slots 5-10 are all in the opposite direction. From
equations (5.4) and (5.5), TLL = 60 even though r ph = 0. Even though
the phase EMF has no flat top, the line-line EMF has a 60 flat top and
the torque is ideally constant.1
All of the above considerations relate to ideal trapezoidal back-EMF
waveforms, which can theoretically be realized only with a rectangular
airgap flux distribution and with the conductors comprising a phasebelt
represented by a uniform current sheet on the surface of a smooth
stator. In other words, fringing and slotting are neglected. In practice
the EMF waveforms are significantly modulated by fringing of the magnet
flux and by slotting, which not only affects the flux distribution but also
concentrates the conductor locations (Chapter 8). Consequently the
torque production is not pcrfect even in the ideal cases.
5.3.2 Delta connection
A careful analysis needs to be made in order to justify the use of a delta
connection based on the number of poles, slots, pole arc and the choice
of the back-EMF wave shape. A particular automatic winding method for
exterior-rotor motors could be so important from a cost point of view as
lo dictate a delta winding because a single needle-winding machine or fly-
winding machine can be programmed to begin with one "start end of
copper wire, wind the first phase, pull a "loop", wind the second phase,
pull a "loop" and wind the third phase. The finish lead is twisted to the
start and the three leads are ready to connect to a circuit board.
The circuit diagram for delta connection of three phase windings is
shown in Fig. 10.2. For squarewave operation the transistors commutate
in the sequence 1,2,3,4,5,6,1... and there are two transistors conducting

It was stated in Ref. [10] that the amperc-conductor distribution in Fig. 5.13*
catmol be realized with a wye-connected winding, but that is only true for a single-layer
winding. The winding in Fig, 5.3 is a double-layer winding.
5-21
D esig n o f bru sh less perm a nent -m a g n et m o t o r s

0 30 SO 90 120 150 180 210 240 270 300 330 360

'//7 7 7 7 7 W )% v y > /7 /7 //, J -J J J

kzzzzzzm

Fig. 5.10a Idealised squarewave motor : delta connection, 60' flat-top phase EMF
waveforms

at any one time, exactly the same as for the wye connection. For sinewave
operation there are generally three transistors conducting at any time.
5-22
5. E l e c t r ic a l D esig n

Fig. 5.10b Idealised squarewave motor : delta connection, 120 flat-top waveforms.
The idea] phase current waveforms are not realisable, because of
distortion by the icro-sequence cuirent.

5-23
D esig n o f bru sh less perm a nent -m a g n et m o t o r s

The lines are labelled A,B,C and the phases are labelled 1,2,3. In the
delta connection the line-line voltages and EMFs are identical to the
phase voltages and EMFs, i.e. AC = vv t^c = v%, and = v3. The line
currents are *a = *1 - *B = *1> an^ *c = *3 " *2-
The delta connection is the dual of the wye connection, in the sense that
relationships between line and phase voltages in the wye connection are
transferred to the currents in the delta connection, and vice versa. A
complete set of waveforms is shown in Fig. 5.10a, assuming phase EMF
waveforms with a 60 flat top. Such waveforms would be produced by the
motor of Fig. 5.136, which has a magnet pole-arc of 120 and the same
stator winding as in Fig. 5.1. Since the machine is delta-connected, the
line-line EMF waveforms are identical to the respective phase EMF
waveforms.
In Fig. 5.10a it appears that there is a 30 phase shift between the
fundamental component of the line current and the fundamental
component of the phase current (e.g., between iA and ij). This is the
same as in the sinewave motor, where ideally the fundamental is the only
non-zero component.
To understand the ideal form of this machine it is necessary to recognize
that throughout each 60 interval the delta is connected between two
lines with the third terminal of the delta open-circuited. For example,
during the 30-90 interval Ql and Q 6 are on, so the delta is connected
as shown in Fig. 5.12, with phases 2 and 3 connected negatively in series
across the supply, and phase 1 connected positively across the supply. If
the motor is "properly designed", then the line current / divides between
the two parallel branches in the ratio 2:1, with (2 /3 )/ going through
phase 1 and (1 /3 )/going negatively through phases 2 and 3.
During the first half of the line A conduction interval, from 30-90,
= 1 unit while ij = 2/3 unit, so the torque contributed by phase 1 is Tj
= fij x = 2/3 unit. During this period is decreasing from its flat-top
value of = -1 unit to zero, while = -1 /3 unit. The average torque
contributed by phase 2 is therefore 1/2 x (-1) x (-1/3) = 1/6 unit
Similarly is increasing from zero to -1 unit and so the torque
contribution from phase 3 averages 1/2 x (-1) x (-1/3) = 1/6 unit, the
same as the average from phase 2. The instantaneous torque
contribution T2 decreases at the same rate as the instantaneous
5-24
5. E l e c t r ic a l D e s ig n
contribution 73 increases, so their sum + Tg remains constant at 1/3
unit. Added to Tlt this produces a constant torque of 1 unit throughout
the 60 interval.
During the second half of the line A conduction interval, from 90-150,
the process is repeated but with all the respective EMF's and currents
commutated to the successive interval: thus el - - eg, *j -e^ - j, etc.
Thus the ideal squarewave delta-connected machine, correcdy
commutated and with the correct EMF and current waveforms, also
produces perfectly constant torque.
The production of constant torque by the delta-connected motor with
60 flat-topped phase EMF's is shown in Fig. 5.136. The magnet has a
120 pole-arc and the phasebelts are 60 wide. Unlike the wye
connection, all three phases conduct simultaneously, even though only
two lines are conducting. Moreover, the ampere-conductors in the three
phasebelts are not all equal. In Fig. 5.136, the instant drawn is at 30 in
Fig. 5.10o: e$ = 0, while el and are both equal to 1 unit. As the rotor
rotates counterclockwise, diminishes as the phasebelts of phase 2 are
"uncovered by the passing magnet, while increases as the phasebelts
of phase 3 are progressively "covered' at the same rate. Meanwhile e]
remains constant at 1 unit as the phasebelts of phase 1 remain covered
throughout the whole 60 interval from 30-90. The current in phase 1
is (2/3)7, and this produces a constant positive torque throughout the
interval. The current in phases 2 and 3 is (-1/3)7, and these phases
produce positive torque contributions decreasing and increasing
respectively at the same rate, as described earlier. The constancy of the
total torque follows from the fact that the effective ampere-conductor
belt width exceeds the magnet pole arc by 60.

The 2:1 division of current in Fig. 5.12 is not as obvious as it looks. If the
phases were pure resistances of equal value, this division of current would
be obvious. But the phases have back-EMF (and leakage inductance
both self and mutual). Under pure DC conditions the inductances have
no effect, but the designer should ensure that the back-EMFs are
correcdy balanced at all times, otherwise, with an imbalance in back-EMF
between the two branches of Fig. 5.12, there will be a net EMF around
the delta and a potentially large current will flow, limited chiefly by the
winding resistances, which are generally small. This current does nothing
useful, but it produces additional I^R losses and torque ripple.

5-25
D e sig n o f b ru sh less perm a n en t -m a gnet m o t o r s

In Fig. 5.10 it can be checked by adding together the waveforms of a ,


&, and ^ are such that their sum, the so-called residual or zero-sequence
EMF e^, is zero at every instant:
e0 = e, + e2 + ez = 0 (5-6)

Fig. 5.10i> shows what happens when a delta connection is used with a
motor having 120 flat-top EMF waveforms. As we have seen, such
waveforms would be produced by the winding of Fig. 5.1 with a magnet
pole-arc of 180. The resulting zero-sequence EMF has a peak value
equal to that of the phase EMF. The phase current waveforms can no
longer be assumed to be of the same ideal form as in Fig. 5.10a, because
the zero-sequence EMF drives a potentially large current around the
delta which increases one of the branch currents in Fig. 5.12 while
decreasing the other one. The resulting current waveform depends on
the size of the zero-sequence current.
Fig. 5.14 shows a computer simulation of the phase current waveform
with a delta-connected motor having a magnet pole-arc of 180 and the
same stator winding as in Fig. 5.1. With no fringing this produces 120
flat-top phase EMFs. This figure is included to show the extreme
distortion of the phase current {mainly third-harmonic) and the
resultant torque ripple, which is very large. The zero-sequence current
dominates the phase current and distorts its waveshape almost beyond
recognition. Since the circulating current is absent from the lines, this
undesirable state of affairs can pass completely undetected. In fact, it
may appear only as an u nexplained" tendency of the motor to overheat
Notwithstanding the problem of circulating currents, many brushless DC
motors are made with delta windings and in some case no precautions
are taken to eliminate the residual EMF. In very small motors the phase
winding resistance may be large enough to limit the circulating current
to a safe value, but it still produces unnecessary losses.

5-26
5. E le c t r ic a l D e sig n

12 '2 3

31 , u
------------- >
u i -------------------------------

Kg. 5,11 Loop currents in wye connection. Each one flows for 60

Fig. 5.12 Division of DC current in della connection

5-27
D esig n o f brushless perm a nent -m a gnet m o t o r s

Fig. 5.13 (a) Squarewave motor with 180 magnet pole-arc and 120 ampere
conductor distribution, (b) Squarewave motor with 120 magnet pole-arc
and 180 ampere-conductor distribution. The phase axes correspond to
those in Figs. 5.1 and 5.3. The numbers indicate the conductors belonging
to phases 1,2.3. Filled circles represent conductors in one direction;
crossed circles represent return conductors in the opposite direction,

The way to suppress the zero-sequence EMF is to eliminate triplen


harmonics from the phase EMFs, that is, harmonics of order Bn, where n
= 1,3,5...
There are basically two ways to do this: one is to eliminate the third
harmonic from the space-distribution of magnet flux, and the other is to
select a winding for which the third harmonic winding factor is zero.
The second method can be thought of as "filtering out" the unwanted
component of the magnet flux distribution.
With no fringing or slotting effects and an ideal rectangular magnet-flux
distribution, the nth harmonic can be eliminated from this flux
distribution by shortening the magnet pole-arc by 1 /n of the pole-pitch.
For example, to eliminate the third harmonic the magnet pole-arc should
be reduced by 1/3 x 180, i.e. from 180 to 120 electrical degrees.

5-28
5. E l e c t r ic a l D e s ig n

Tig. 5.14 Computer simulation of delta-connected motor with 180 magnet pole-arc
and near-I20 flat-top phase EMF waveforms, showing the large zero-
sequence EMF around the delta and the consequent distortion of the
phase currents and the torque ripple. The line currents are 120
squarewaves, and the zero-sequence current docs not appear in the lines
or in the controller: consequently it may pass undetected. The winding is
the same as in Fig. 5.1.

The mathematical statement of this requirement is that sin (npM/2 ) =


0. In this case P M/2 = 180 so the condition is clearly satisfied, not
only for the third but also for all the triplen harmonics. Unfortunately,
because of fringing and slotting effects, it is impossible in practice to
eliminate the third-harmonic component of the flux distribution exactly.
Thr alternative way to suppress the zero-sequence EMF is to select a
winding with a third-harmonic winding factor of zero, i.e., ^ = 0
(Chapter 6). The simplest winding with this property is a "chorded"
winding with one coil per pole having a coil pitch yc = 2/3 (120), and
no skew. The harmonic winding factor is then equal to the pitch factor:
1.e. = = sin (3Yc'n/2) = 0. There are several different winding
techniques that can be used to achieve this result.

5-29
D esig n or bru sh less perm a nent -m a gnet m o t o r s

[Aaptf] \ ft.B_lB.BCC
1.80 * 1.0*1
PHArf onlu clbotm ROTDS PCflTYION

Phase cu rran t 1-j


0.90
1---------
I
-u.nc
-l.flO
[Uolur) a 1.0.0
i E-*.r. v-
ooroo DtornoN . a.bjb.buc
2,00 ______________ ^ ^ ^ . P h a s e E M F e*|

^ S"
^ \ /
-2.00 Z e ro -s e q u e n c e vo ltag e e 0 V ----------------------- /
C-N TOBflUE Tt fiOTOfl POSITION, A,t 15.BDC
2.60
[N.l K 1.0.-I

2.00
I.SO Torque
1.00
0.60
0. 00 0.39 O.tfO 0.00 1.20 I.BO 1.00 2.10 2.AQ 2.70 3.00 3.30 3-<0
Pto>- position [*lc dq' x 1.0*2

f .b.5 .15 Computer simulation of delta-connected motor with 180 magnet pole-
arc, with the winding of Fig. 5.3. The phase EMF has a narrower flat top
and the zero-sequencc EMF ii eliminated by the third-harmonic winding
factor.
Interestingly enough, the winding of Fig. 5.3 also has k = 0 provided
that all coils have equal numbers of turns, even though it has two
separate coil-pitches neither of which is 2/3. In effect, the outer coils 1
and 2 are "under-chordcd" relative to the ideal 2/3 pitch, while the inner
coils 3 and 4 are "over-chorded"; in the superposition of EMF's the
triplen harmonics induced in these coil-pairs cancel, provided that they
are connected in series.
Fig. 5.15 shows a simulation of the motor of Fig. 5.3 with delta-connected
windings, 180 magnet pole-arc, and the double-layer winding. It can be
seen that the zero-sequence loop EMF is zero and so is the torque
ripple.
The designer thus has an additional task when designing motors for delta
connection, namely to ensure that the zero-sequence EMF around the
delta is zero. The PC-BDC. program plots the waveform of the loop EMF
Kq = + <2 + ^ along with the phase and line-line EMFs when the
5-30
5. E l e c t r ic a l D e s ig n
Ending connecdon is delta, Figs. 5.14 and 5.15. Any departure from zero
shows up immediately and can be corrected by the necessary adjustment
ro the parameters mentioned. In sinewave motors, the same requirement
exists. In wye-connected motors the problem does not arise because the
zero-sequence current is forcibly suppressed by the three-wire connection
the neutral point of the windings, and there is only one current path.
5,3.3 Flux/pole and magnet utilization
Itis helpful to construct a model of Figs. 5.13a and b using construction
paper, in which the three separate rings, suitably shaded or coloured, are
cut out andpinned at the centre so that the two inner ones can be
independently rotated while the outer phasebelt ring remains stationary.
This model will prove invaluable in sorting out the sequence of events as
the rotor rotates.
While the flux-distribution of the magnet rotates with the rotor in a
continuous fashion, the MMF distribution of the stator remains stationary
for 60 and then jumps to a position 60 ahead. This motor is not a
rotating-field machine in the sense associated with AC machines.
In Fig. 5.13a, the production of smooth, ripple-free torque depends on
the fact that the magnet pole arc exceeds the MMF arc by 60. The
magnet is therefore able to rotate 60 with no change in the flux-density
under either of the conducting phasebelts. An inevitable result of this is
(hat only 2/3 of the magnet and 2/3 of the stator conductors are active
at any instant, although all of the stator ampereconductors are active.
Similarly in Fig. 5.13b the whole of the magnet is overlapped by
conducting phasebelt, but 60 of ampere-conductors are wasted, either
waiting for the magnet to arrive, or waiting for the next commutation
after it has passed. The motor of Fig. 5.13ft may therefore appear to have
higher per-unit copper losses than that of Fig. 5.13a. Offsetting this
disadvantage is the fact that for the same magnet flux-density, the
flux/pole in Fig. 5.136 is only 2/3 that in Fig. 5.13a, so that only 2/3 of
ihe stator yoke thickness is required. If the stator outside diameter is kept
the same, the slots can be made deeper so that the loss of
ampere-conductors can be at least partially recovered. Consequently the
efficiency of the motor of Fig. 5.136 may not be very much lower than
that of Fig. 5.13a, provided that the extra copper is used.

5-31
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

The selection of magnet pole-arc and winding layout is subject to many


other constraints besides those discussed here. For example, changing
the magnet pole-arc changes the flux and flux-densities throughout the
magnetic circuit, while any change in the winding layout can affect the
winding resistance and the ease of manufacture. Not only the
instantaneous torque but also the average torque is a function of all these
variables, as is the efficiency. The totality of the designer's job is
potentially very complicated, and can be gready assisted by
computerization.

5.4 Unipolar 3-phase connection


The unipolar 3-phase connection is shown in Fig. 10.12, and the ideal
waveforms in Fig. 10.13. This connection is sometimes used in order to
save half the number of transistors, but it is less efficient than the normal
wye and delta arrangements. The commutation of this circuit is analysed
in section 10.5. With only three power transistors, it certainly has an
appeal for very low cost applications.

5.5 2-phase and single-phase connections


Fig. 5.16 shows a single phase winding controlled by a full-bridge PWM
inverter. A 2-phase system can be made up of two of these circuits. The
axes of the phase windings in the motor are displaced from each other
by 90 electrical degrees. With this controller circuit there is no
distinction between line and phase quantities. The currents required to
maintain continuous rotation in the steady state are shown in Fig. 5.17
as 90 squarewaves, and ideally each phase EMF has a trapezoidal
waveshape which is in phase with the corresponding current waveform.
The resulting torque is ideally constant, and can be controlled by
chopping or current regulation, which is shown in Fig. 5.17 and
discussed in Chapter 2.
When used with two-phase motors the controller in Fig. 5.16 requires
eight transistors, two more than are required in a three-phase drive. It
is possible to replace one phaseleg (e.g. transistors Q3 and Q2 in Fig.
5.16) by capacitors that split the DC supply voltage into two equal levels,
so reducing the number of power transistors from 8 to 4. However, this
means that the motor windings must be re-wound for half the voltage
5-32
5. E l e c t r ic a l D e s ig n

Fig. S.lfi Electrical circuit, 1-phaje and 2-phase controllers

and twice the current, and the remaining power transistors must have
double the current rating. So it is unlikely that there will be any major
saving in total Silicon area. At the same time the bus-splitting capacitors
add cost and bulk and may limit the performance in other ways.
Moreover, there is less freedom to design a two-phase motor with
sufficient overlap or even any overlap between the flat tops of
successive phase EMFs, when compared with the three-phase motor.
These considerations may help to explain why the two-phase brushless
DC motor whether squarewave or sinewave is rare, although it is in
production by a number of companies.
One of the concerns with the two-phase brushless motor is the ability to
start from any rotor position. This is of less concern with sinewave
motors than with squarewave motors. However, with single-phase motors
ilxs a fundamental problem becaue, with symmetrical classical designs of
The type discussed so far in this chapter, the torque over half of each
electrical revolution would be zero or negative. In order to produce
positive torque over the entire revolution, even-order torque harmonics
jttiy be introduced by the addition of saliencies as described by Kenjo
<ind Nagamori [ 1].

5-33
D e s ig n o f b r u sh l e s s p e r m a n en t -m a g n e t m o t o r s

0 30 60 90 120 150 180 210 240 270 300 330 360 Elec ctogrwa

-30
Z T
'2

Q1 Q3 Q1
I
Q2 Q4 Q2

Q1 Q3
02] Q4

ISP Q1 duty-cycled

Chopping transistor

Is p - D4 duty-cycle 1-d

Chopping diode

ISP Q2

Commutating transistor

D3

Commutating diode

Fig. 5.17 Current and EMF waveforms 2-phase


5-34
5. E l e c t r ic a l D esig n

5.6 The EMF constant ^


The peak (flat-topped) value of the phase EMF can be calculated using
the waveforms in Fig. 5.4, assuming a magnet pole-arc of 180. We have
already established that if all the coils have the same number of turns Nc,
then the peak EMF is the same in each coil regardless of its pitch.
From equation (5.1) the back-EMF cl in coil 1 is given by the slope of
the curve of <|/cl(), multiplied by (i>m. The slope of '|'cl() is equal to
V-s per electrical radian, where i)M is the peak flux-linkage of a
futt-pitch coilaligned with the <i-axis. If there are more than two poles,
then dty/dX, = V-s per mechanical radian, where pis the number
of pole-pairs. Evidently, with a rectangular magnet flux distribution,

^ <5 7 >
where <J> is the airgap flux due to the magnet, B is the airgap flux-demtfy
due to the magnet, D is the stator bore, and L is the stack length (active
length). Hence
ec = NcB umDL (5 .8 )

In [V]. If all the coils in one phase have the same back-EMF induced in
them, in phase with each other, the subscript 1 is unnecessary, as in
equation (5.8), and the total phase EMF has a peak value eph equal to ec
multiplied by the number of coils in scries per phase. If the number of
turns in series per phase is N^h, then
^ = N^B<*mDL (5 .9 )

If Zis the total number of conductors in the machine, and the number of
parallel paths is a, then = 1/3 x Z/2a.
In squarewave 3-phase motors, if the windings are wye-connected then
the line-line back-EMF is = 2e h. In delta-connected windings =
This is the back-EMF "seen" by the controller.
The EMF constant is defined as Thus, using equation (5.9),

5-35
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

(5.10)
in Volts per (rad/s). Alternative expressions for Aj. are developed in
Chapter 7, which also discusses methods of measurement.
The EMF ec in a full-pitch coil can also be thought of as twice the EMF
in each of its two conductors, or 2BLv, where v is the linear velocity of
the magnetic field past the conductor, v = w^D/2. This expression can
easily be used to develop the back-EMF formulas discussed hitherto, and
it is tempting to use it, but it is subject to a number of uncertainties. The
BLv concept of back-EMF is that of a conductor moving through a
magnetic field, as in every physics text-book, but in most electric motors
the conductors are not located in the flux at all: they are located in slots,
and for all practical purposes the permeability of the teeth is so high that
no flux passes through the conductors at all! To overcome this difficulty
it is assumed that the conductors are equivalent to filaments at the
diameter D, located at the slot centres, or to uniformly distributed
current sheets on a smooth stator surface. But this is artificial and is not
a satisfactory assumption when the number of slots/pole is small or
fractional, and especially when the slot-openings are large (see Chapter
8). It is for these reasons that the EMF equations are derived here
directly from Faradays Law (equation (5.1)). This is a more modem
approach and it is more rigorous because the concept of flux-linkage is
more rigorous than the concept of flux-cutting by conductors which are
not even in the magnetic field.

5.7 The torque constant kj


5.7.1 Basis of torque productioncoenergy
The torque constant is the most fundamental parameter in brushless
servomotors and is basically the torque per ampere of DC supply current,
although minor variations of definition are possible for sinewave and
squarewave motors (Chapter 8).
The most fundamental equation for the torque of an electrical machine
is
d \V (5.11)

5-36
5. E l e c t r ic a l D e s ig n

FLUX-LINKAGE

Ktg. 5.18 Idealised flux-linkage vs. position characteristics with impressed constant
current, showing the production of torque by incremental changes in co-
energy

where is the rotor position, i is the current, and W is the coenergy.


The coenergy is defined as the difference between the product i|ri and
ihe stored field energy. In brushless motors the stored field energy is very
small and the coenergy is nearly equal to the product tjri, and in Fig. 5.18
it is represented by the area beneath the horizontal line representing the
flux-linkage at a particular rotor position , bounded on the right by the
vertical line at the current I. In the squarewave wye-connected motor
there is only one current and it is not necessary to treat i as a vector of
currents. The flux-linkage is the flux-linkage of two phases in series
(assuming wye connection).
When the rotor moves through an infinitesimally small angle if the
current remains constant the coenergy changes by the amount AW'
shown in Fig. 5.18, and equation (5.11) gives the electromagnetic torque:
T = A . (5-12)

5-37
D esig n o f bru sh less perm a nent -m a gnet m o t o r s

We have already been using diji/d in equation (5.1) and in the previous
section to define kE, and it follows immediately that kT = k^. Further
discussion and analysis of the torque constant and the EMF constant, and
their measurement, are given in Chapters 7 and 8.
5.7.2 Torque linearity
It is central to the design of servomechanisms that the torque of a
brushless servo motor is proportional to the current, with a fixed1
constant of proportionality kj. A graph of torque vs. current is a straight
line, k^I. In practice the torque/current graph may deviate slightly
from a straight line, mainly because of saturation of the stator teeth and
other parts of the magnetic circuit. The torque linearity is defined as the
ratio of the actual torque to the value of k jl at a given current I, which
is typically quoted as 100% of rated current.
5.7.3 Demagnetization
The magnets can be partially demagnetized by overcurrent or
overtemperature, or a combination of both. Brushless motors should be
specified to suffer no more than a certain percentage demagnetization
at a certain current and temperature, for example, 5% at 300% rated
current at a winding temperature of 155C. The 5% refers to the
permanent decrease in the value of fej- (or k^) following a test at this
level.
Test procedures should ensure that the demagnetizing current is applied
at all orientations so that the worst case is definitely covered. The
calculation of partial demagnetization effects is difficult by manual
calculation methods and is best tackled with finite-element or boundary-
element methods.

5.8 Calculating the number of turns


In some cases the torque constant may be specified by the customer, and
it is often quoted in catalog data. In that case, once the basic dimensions
of the motor and the type of winding and its connection are decided,
equation (5.10) can be used directly to calculate the number of turns in
series per phase.

5-38
5. E l e c t r ic a l D e s ig n
The total number of conductors and parallel paths, and the number of
strands-in-hand in the winding of each conductor, must be adjusted to
m atch the available supply voltage while maximizing the slot-fill factor to
keep the winding resistance and copper losses as low as possible.
If the torque constant is not specified, the operating speed of the motor
certainly will be. Maximum speed will usually correspond to operation
with maximum voltage, i.e. with no chopping. In most cases the
maximum speed will be only slightly less than the no-load speed because
the slope of the speed/torque characteristic is generally quite shallow
because it is proportional to the winding resistance, which is made as
small as possible in the interests of high efficiency. This means that the
no-load speed JV0 can be taken as representative of normal operating
speeds. In mechanical rad/s, the no-load speed is given by equation
(1.7), so in rev/min
N0 = x J = J l * L. (5.13)
0 ^ 2-n- kE 2-rr
From this equation, if N0 and the DC supply voltage Vs are known, AE
can be calculated and then the number of turns in series per phase can
be calculated from equation (5 .10).
This process is simple and is adequate for many purposes, but it is
sometimes necessary to refine the estimate of the required number of
turns particularly if the design constraints in the specification are very
tight. For example, in low-voltage systems (12V or 24V) the number of
turns can be critical in determining the correct motor operation, and
because it has a bearing on the required current, it also affects the
choice and rating of power transistors.
For sinewave motors the number of turns can be calculated in the same
way, but the formula for is different: see sections 7.4.2 and 7.4.3.

5.9 Winding inductances and armature reaction


5.9.1 Importance of inductance
In Chapter 4 the effect of phase current on the magnetic flux was
discussed, especially in relation to the possibility of demagnetizing the
magnets. The "armature reaction" effect was viewed as a distributed
5-39
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

field effect, with the possibility of bulk or local demagnetization.


"Armature reaction" also has important effects in the electric-circuit
operation of the brushless motor and its controller. The voltage supplied
to the phases is in the form of chopped or pulsed DC. The phase
circuits contain resistance, back-EMF, and inductance.

Fig. 5.19 Effect of inductance on current waveform at high speed


The current waveform is, in effect, the response of the electrical circuit
to the series of voltage pulses. As in any electrical circuit, the inductance
plays an important part in determining the rate of rise or fall of current.
At high speed, the inductance may limit the rate of rise of current so
much that the desired current is never reached.
5-40
5. E l e c t r ic a l D esig n
The effect of high speed on the current waveform is illustrated in Fig.
5.19. When the supply is switched to the phase winding, the net driving
voltage is Vt - E, and the rate of rise of current is

= Vs I E (5.14)
dt L
where L is the circuit inductance. For example, in a wye-conncctcd
squarewave motor, L is the inductance of two phases in series. At low
speed the back-EMF E is much less than the supply voltage Vs, so the rate
of rise of current is large. This is shown in the left-hand diagrams in Fig.
5.19a. The current rises so quickly to the set-point value 7jp, that it has
to be limited by chopping the supply voltage.
At high speed the net driving voltage Vs - E is reduced because E is
closer to Vs. Consequently the rate of rise of current is lower, and at a
sufficiently high speed the current fails to reach the set-point value
during the available conduction period. Clearly the average current is
reduced, and so the torque is reduced, as shown in Fig. 5.20.
Matters are somewhat worse than indicated by equation (5.14) because
the current is really intended to rise to the set-point value within a
certain angle of rotation, not a certain time-interval (see bottom diagram
in Fig. 5.19A). As the speed increases, the angle of rotation during a
given time interval increases proportionately, or, put the other way
round, the time interval available to build the current decreases in
inverse proportion to the speed. Therefore, in a fixed angle of rotation
at high speed, the current rises to a lower value than it would at low
speed, even if the net driving voltage was the same.
These two factors combine to cause the torque to collapse as the speed
is raised above a certain level. Mathematically, the effect can be
expressed by the rate of rise of current with respect to rotor position,
rather than with respect to time:

A = dI>dt = Vs ~ E = (5.15)
dO d d /d t w mL a) m L

This equation clearly shows the double effect of the increasing

5-41
D e s ig n o f b r u sh l e s s p e r m a n en t -m a g n e t m o t o r s

speedonce through the back-EMF constant and once through the


inductance. For high-speed operation of course both and L can be
reduced by decreasing the number of turns (at the expense of the torque
constant fcj-).

Fig. 5.20 Effect of inductance on speed/torque characteristic at high speed


5.9.2 Inductan ce components
The phases have self-inductance and mutual inductance with the other
phases. There are three main components of the self-inductance and
three main components of the mutual inductance. These are
1. the airgap component,
2. the slot-Uakage component, and
3. the end-urinding component.
These are dealt with separately in the sections that follow. They
correspond to different portions of the flux linking the windings. The
division of the flux into separate portions is a somewhat arbitrary
concept, based on equivalent magnetic circuits. It is a very old concept
that has been used in machine design for decadcs, because it is highly
amenable to manual calculation or rapid computer calculation. The
most accurate mathematical modern methods for calculating inductance
are numerical methods (such as the finite-element or boundary-element

5-42
5. E l e c t r ic a l D e s ig n
methods), which solve the entire field problem and do not distinguish
between "airgap flux" and "slot-1eakage flux". These methods give the
total flux-linkage of a winding, from which the self- and mutual
inductances can be obtained by dividing by the respective currents.
Alternatively, they compute the total stored energy in the magnetic field.
This energy is then expressed in the form
fKF = -2 L I 2 (5.16)
where I is the total current at the terminals. The inductance L is
extracted by re-arranging this equation. An alternative method is to
evaluate the flux-linkage of the winding as an integral of the vector
potential A along the length of the conductors:
(5.17)
From a circuit-analysis point of view, it is important to recognize the
separate effects of self-and mutual inductance. For example, in the
three-phase squarewave motor with two phases on, the inductance
between two line terminals of the motor (say, A and B) is
(5.18)
where LA and are the self-inductances of phases A and B, and is
the mutual inductance. The positive sign applies if the two phases cany
current in such a direction that their fluxes arc essentially additive, as in
Fig. 5.21a. The negative sign applies if they are in opposition. It is
possible for two phases to have very low or even zero mutual inductance,
if their axes are orthogonal (Fig. 5.21ft), as in the two-phase motor.

(a) (b)
fig 5.21 Total self-inductance of two coils connected in series with mutual
coupling, (a) Fluxes are essentially additive, so L - LA + Lp + L ^ . (b)
Winding axes at 90*, giving zero airgap mutual inductance.

5-43
D esig n o f bru sh less perm a nent -m a g n et m o t o r s

It is perhaps worth spelling out in more detail why the fluxes are
essentially additive in Fig. 5.21a. The dot convention means that when
positive current is applied, the flux is in the positive direction; and
positive current is that which enters at the dotted end of the coil. In a
wye-connected three-phase motor (with no neutral connection), positive
current entering phase A is negative current leaving phase B.
Accordingly, the flux produced by phase A is in the positive direction
(shown leaving the dotted end), while the flux produced by phase B is in
the negative direction (shown entering the dotted end). The winding axes
are not aligned, and therefore the fluxes do not add directly. If the
winding axes are at 90 there is no common flux-linkage and no mutual
inductance. If the angle between the winding axes were less than 90
the fluxes would be essentially in opposition and the negative sign would
be required in equation (5.18).
5.9.3 Airgap self inductance of single coil
The basis forcalculating the airgap self-inductance L of a single coil is
shown for a full-pitch coil in Fig. 5.22. Thisshows the magnetic flux
established by a full-pitch winding with one slot per pole per phase. The
total MMF around a complete loop or flux-line is equal to Nci, where Nc
is the number of conductors in the slot and i is the current. Nc is also
the number of turns in the coil.
If the steel in the rotor and stator is assumed to be infinitely permeable,
then the MMF is concentrated entirely across the two airgaps. Across
each airgap the MMF drop is Nci/2. If the flux is assumed to be radial in
the gap, the magnetizing force in each gap is

H - 2sL. (5-19)
2/
In a surface-magnet motor the gap g " includes the radial thickness of
the magnet as well as the physical airgap g, which may be modified to gl
= Kcg by the Carter coefficient Kc for the stator slotting. A reasonable
approximation for g " is then
g " = g< + J . (5.20)
Mrec

5-44
5. E l e c t r jc a l D e sig n

% 6.22 Calculation of airgap self-inductance of a single full-pitch coil

The flux-density produced by this magnetizing force at the bore of the


slator is
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

The subscript a is added to emphasize that the flux is that which is


excited by armature reaction, i.e., by the stator current. The flux
distribution around the airgap is plotted in Fig. 5.22. The flux-linkage of
the coil is
i|r = (5.22)

and the airgap self-inductance is


" 'M o^Adc'i (5.23)
8 * 2 g"
in [H]. This result has been derived for a 2-pole machine. If there are
p pole-pairs and if Nph is the number of turns in series per phase, then
I - Jfc - (5 24)
g ' j ~ I p 'g "
If there are a parallel paths through the winding, thisformula is
unchanged provided that Afph is interpreted correctly as Nc x 2p/a. An
example is shown in Fig. 4.10, with p = 2 (i.e., 4 poles) and one coil per
pole. In Fig. 4.10c, all the coilsare in series and a = 1, sothat = Ne
x 2 x 2/1 = 4Nc. In Fig- 4.106, there are two parallel paths, a = 2, and
JVph = Nc x 2 x 2/2 = 2JVC. The total number of turns in the phase
winding is 2pNc, and if they are reconnected in a parallel paths the
inductance is automatically reduced by the factor a , because A^h is
inversely proportional to a.
5.9.4 Airgap mutual inductance between two coils
The mutual inductance between two phases whose axes are displaced by
120 electrical degrees is important. This can be calculated by adding up
the flux-linkage of a second coil placed in the field of the first one. For
a second coil with conductors located at the angles 90 + 120 = 210 and
-90 + 120 = 30 the flux-linkage is

'I'll =
(lrr 7T1 ' 7r (5.25)
; . 6 2 , + 2 " eJ

The positive and negative signs account for the direction of the

5-46
5. E l e c t r ic a l D esig n
flux-density in the airgap. Substituting the expression for from
above, we get
A. _ _____________________________
8 h 3 2g

in [H], Note that


= - I. (5.27)
h
It is interesting to contrast this ratio with the value -1 /2 which is
obtained with sine-distributed windings (Chapter 6).
Once L has been calculated, equation (5.27) can be used for A f,
provided that the flux distribution is a squarewave as shown in Fig. 5.22.
Physically this result makes sense : if the second coil was aligned with the
first,the self and mutual would be equal. If the second coil was displaced
90 from the first, the mutual would be zero. With a rectangular flux
distribution produced by current in the first coil, and with concentrated
windings of one slot per pole per phase, the flux-linkage of the second
one varies linearly with the angle between their axes. By rotating the
second coil to a position one-third of a right-angle past the 90 position,
it picks up or links one-third of the flux, but in the negative direction
relative to its own positive axis.
If there are p pole-pairs with all turns in series, then Nc is substituted by
'Vph = N c / p an d
M _ 2 7r lI0AU 2A*lr/i (5.28)
8 3 2p2g"
If the total number of turns is unchanged and the coils are reconnected
in a parallel paths, then both the self and mutual inductances are
automatically reduced by the factor a2 because Nph is inversely
proportional to a.

547
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

Fig. 5.23 Calculation of airgap self-inductance or a winding with 2 slou/pole

5.9.5 Examples of inductance calculations


Consider a w in d in g with 2 slots per pole per phase, as shown in Fig. 5.23.
The conductors o f a second, unexcited phase are shown in their correct
positions for calculating the mutual inductance between phases. By *
process similar to the one above, in which the flux-linkages of the
individual coils arc added together, the self inductance is shown to be
5-48
5. E l e c t r ic a l D e sig n

L. (5.29)

where
(5.30)
is the winding factor for q = 2 slots per pole per phase, i.e. 0.833. The
self-inductance is thus only 83.3% of the value which would be obtained
with the same number of turns per phase concentrated in one slot per
pole per phase. This concurs with the general rule that the inductance
of a coil is increased when its conductors are concentrated, together.
When the mutual inductance is evaluated, using the same method as
before, it is found that the distribution of the second winding cancels the
effect of the step in the flux distribution, so that the actual value of the
mutual inductance is the same as with one slot per pole per phase
(provided the total turns are the same). The ratio between the self and
mutual inductances is therefore

(5.31)

Once Lnhas been


1 the *. .calculated,
. is> asthis
< equation
. can
rFig. w nbe used for Mo provided
that flux distribution shown in 5.23.
A further example is shown in Fig. 5.24 with 3 slots per pole per phase.
The conductors of a second, unexcited phase are shown in their correct
positions for calculating the mutual inductance between phases. By a
process similar to the one above, in which the flux-linkages of the
individual coils are added together, the self inductance is shown to be
given by equation (5.29) where
16 (5.32)
K 27 q
is the winding factor for q = 3, i.e. 0.802. The airgap self-inductance is
thus only 80.2% of the value which would be obtained with the same
number of turns per phase concentrated in one slot per pole per phase.

5-49
D esig n o f bru sh less perm a nent -m a gnet m o t o r s

Fig. 5.24 Calculation of airgap jelf-inductance of a winding with S slots/pole

It is again found that the distribution of the second winding cancels the
effect of the steps in the flux distribution, so that the mutual inductance
has the same value as with one slot per pole per phase (provided the
total turns are the same). The ratio between and is given again by
equation 5.81, and its value is -0.415. Once has been calculated, (his
equation can be used for provided that the flux distribution is as
shown in Fig. 5.24.
5-50
5. E l e c t r ic a l D esig n

In field

fig, 5.25 Geometry of a jingle stator coil and associated flux-paths. The coilside
locations are defined by the angles 0G for the "go" conductor and 0R for
the "return" conductor.
5.9.6 General case of airgap inductance
The airgap self-inductance of any distribution of conductors can be
calculated automatically by computer, following the principles described
in the previous sections, and this calculation can be extended to include
the mutual inductance between any two windings regardless of the
distribution of their conductors, provided these arc known. The
accumulation of flux-linkage in the winding is done by exciting each coil
in turn, and adding the flux-linkage of every coil including the excited
coil. In this way the total inductance can be seen as being composed of
a sa of n self-inductances and n (n -l) mutual inductances, where nis the
number of coils in the winding. The formulation of this process is
ttmple in electromagnetic terms, although it requires careful
programming. The PC-BDC computer program uses a coil table to
manage the inductance calculation. The coil table is a list of coils, each
identified by its number, the number of its turns, and the slot-numbers
of the "go" and "return" conductors.
The basis of the method is shown in Fig. 5.25, which shows a single coil
with the "go" conductor located at 0G and the "return" conductor at 0R.
Applying Amperes and Gauss laws,
5-51
D esig n o f brushless perm a nent -m a cn et m o t o r s

(H i * H 0 ) g " - N cj (5-33)
and
4>i = 0O (5-34)

where H-t is the magnetic field strength in the airgap "inside" the coil
(i.e., between 0G and 0R) and H0 is the magnetic field strength in the
airgap "outside" the coil, g" is the effective airgap which can be
approximated by g-+ fci/l-irec in surface-magnet motors. These sections
of the magnetic field are labelled "infield" and "outfield" in Fig. 5.25.
<t>; is the flux crossing the airgap in the "infield", "inside" the coil, and
is the flux crossing the airgap in the "outfield", "outside" the coil. Now

<f>0 = i[2ir - (9 r - 90)]. ' '


= o - ^ A a k - ' i ^ R 3 1 1 (1 ( 5 3 5 1

If we write a = (0R - QG)/n and use B - \lqH, we get

3 ' ' f ) ; B ' <536)


where the negative sign added to BQ signifies that when isradially
inward, Bq is radially outward and vice versa. The flux-linkage is
* 1. =
(5.37)
= *
so that
v 4- - 2 ( - f ) (W8)

where
r _ irl*ojVc2 /'stk'ri C5.391
-^m ax "
--------- v
2g '

5-52
5. E l e c t r ic a l D esig n

G1

R2

Fig. 5.26 Geometry of jtator coils for the calculation of airgap mutual inductance.

Evidently the inductance is zero if a = 0 or a = 2: in both these cases,


the coil-span is zero. The airgap inductance is maximum if a = 1, which
corresponds to the full-pitch coil, equation (5.23) and Fig. 5.21.
The airgap mutual inductance can be formulated in a similar manner
except that different categories of overlap must be taken into account.
Fig. 5.26 shows the basic geometry and Fig. 5.27 shows the idealised flux-
distribution with partial overlap between coils 1 and 2, assuming coil 1
is the excited coil. The flux linkage of coil 2 due to coil 1 in this case is
^21 ^ c 2 ^ s tk r i +

- - *C2) - *o( 0R2 " *Rt>] (5.40)

where 0G1 is the location of the "go" conductor of coil 1 and 0R1 is the
location of its "return" conductor, and similarly for coil 2. The mutual
airgap inductance between coil 2 and coil 1 is evaluated as
5-53
D esig n o f bru sh less perm a nent -m agnet m o t o r s

Fig. 5.27 Idealised flux distribution for the calculation of airgap mutual inductance
between the two stator coils of Fig. 5.2fi.
A number of different possibilities arise, depending on the nature of the
overlap between the coils. If the conductors of coil 1, the excited coil,
are represented by [ ], and if the "go and "return" conductors of coil
2 are represented by G and R respectively, then four basic categories of
overlap can be identified as
(i) []GR No overlaprGo 2 and Return 2 both outside coil I
(ii) [G ]R Go 2 within coil 1; Return 2 outside
(iii) G[ R] Return 2 within coil 1; Go 2 outside
(iv) [G R] Go 2 and Return 2 both within coil 1
The equations for the flux-linkage of coil m due to current in coil
n, are derived in the same way as equadon (5.40), which is the correct
formula for case (ii), [ G ] R. The total airgap self flux-linkage of the
phase winding is then

(5.41)
tn-l n-\
The mutual inductance between two phases is evaluated in the same way,
cxcept that the summation over m is taken for the coils of one phase
5-54
5. E l e c t r ic a l D esig n
while the summation over n is taken for the coils of the second phase.
5.9.7 Slot leakage inductance : self and mutual
The slot-leakage component of the phase self-inductance and of the
phase-phase mutual inductance is typically of the same order of
magnitude as the airgap component. It is evaluated in the same way as
for induction motors, except that there is an additional component of
slot-leakage flux due to fringing between the tops of adjacent teeth. This
fringing flux can be considered a space-harmonic addition to the airgap
component of armature-reaction flux. In the induction motor this
component of flux would cross the airgap and flow along the rotor
surface without linking the rotor flux, and it is then known as zig-zag
leakage. In the surface-magnet brushless motor the magnet is generally
so thick that there is no zig-zag effect as such.
The procedure for calculating slot-leakage inductance follows that used
for induction motors, with a few differences in technique in order to
adapt the calculation for automatic computation. The procedure is first
to determine the slot permeance coefficient appropriate to the various
combinations of conductors that may be found in one slot (including the
case where there is only one conductor per slot); and then to add up the
contributions to the total phase flux-linkage from every slot, taking into
account the distribution of conductors from the separate phases.
The slot permeance coefficient can be understood by considering the
inductance of a bunch of N conductors occupying a rectangular stator
slot, as in Fig. 5.28. The slot-leakage flux follows the paths illustrated by
dotted lines. Assuming that the steel is infinitely permeable, the main
reluctance is the cross-slot reluctance, which evidently depends on h and
w, the depth and width of the slot Intuitively we can see that if the ratio
w/h is small, the cross-slot reluctance will be small and the cross-slot
permeance and the inductance will be high. The total flux-linkage of the
conductor in Fig. 5.28 is calculated as follows.
The cross-slot flux-density is given by
B = p.H = * -N I. (5-42)
w h
The flux in the elemental tube shown in Fig. 5.28 is <&)>= RLaVdx and this
flux is linked by a fraction x/h of the total number N of conductors,
5-55
D esig n o f brushless perm a nent -m agnet m o t o r s

Fig. 5.28 Calculation of slot permeance coefficient Tor open rectangular slot with
uniform ampere-conductor distribution throughout the slot.

giving rise to a contribution to the total flux-linkage. The total flux-


linkage is therefore

0 hw 0 (5.43)

so the slot inductance is Ls = (i0Lstk7 r x h/Bw. The coefficient h/Stu is


the effective depth /width ratio of the slot, and is known as the slot
permeance coefficient,
The integral in equation (5.43) can be calculated for various distributions
of the conductors, and Fig. 5.29 summarizes some interesting special
cases for the rectangular open slot. Case a is the one just analyzed. Ill
case b all the conductors are concentrated in one filament and P.
h/w. This represents the highest possible inductance. In case c the
conductors produce very little flux other than fringing flux around (lie
top of the slot, and link none of it, so ^jlot - 0 and the inductance Is
virtually zero. In case ^ ^slot = 2/3 x h/w. This is 2/3 the maximum
theoretical value of case b, but it is twice the value obtained in case a with
uniform distribution. In case e, ^slot = 1/6 x h/w, giving only one-half tbe
inductance obtained with uniform conductor distribution, and only one-

5-56
5. E l e c t r ic a l D e s ig n

(a) (b) (c)

(d) (e) (f)

fif. 5.29 Slot leakage inductance: effect of conductor location within the slot.

fourth the inductance obtained by concentrating the the conductors in


the lower half of the slot.
When two conductors share the same slot, one in the top and one in the
bottom, the mutual slot inductance is evaluated using = H0A.tkiV x
h/4-tn, i.e. the slot permeance coefficient for mutual inductance is A/4w.
The rectangular open slot is a rarity in brushless DC motors, partly
because this type of slot is not ideal for automatic winding and partly
because open slots aggravate the problem of cogging torque. A more
common slot type is shown in Fig. 5.30 where the tooth "tangs or
"overhangs" reduce the effective slot-opening without materially reducing
the slot area. In brushless DC motors the rectangular slot is often used
in preference to the round-bottomed slot common in mass-produced
induction motors. Probably the reason is that the rectangular slot has a
greater winding area, which is important to minimize the temperature
rije (brushless motors are often totally enclosed). Induction motors, on
the other hand, tend to be produced in large numbers at minimum cost,
F^nrl the round-bottom slot prolongs the life of the stamping die.
5-57
D e sig n o f brushless perm a n en t -m a cn et m o t o r s

Fig. 5.30 "Rectangular" slots common in brushlcu PM motors

The tooth overhangs augment the slot permeance, making it necessary


to add a separate component t/w0 to ^slot- This helps to "wash out" the
uncertainty in calculating Pslot which arises from not knowing the precise
locadon of the conductors within the slot. If there is one coilside per
slot the slot permeance coefficient should be taken as
1 h
rslot 3w (5.44)
where k is equal to the total slot depth minus the depth t of the tang.
Many detailed formulas have been developed for slot permeance
coefficients of various slot shapes, including allowances for gaps between
coilsides sharing the slots: see, for example, Alger [2] and Veinott [3].
Alger in particular incorporates into the total "primary leakage reactance"
the mutual inductance between coilsides of different phases sharing the
same slot, and he gives formulas for the resulting slot permeance
coefficient which depend on the winding pitch and the type of winding.
This helps in representing the induction motor by a per-phase equivalent
circuit in which the mutual coupling between phases is implicit in the
per-phase self-inductances. However, this model depends on having
balanced three-phase sinewave currents and fluxes, which generally is not
the case in the brushless PM motor. In any case, for computer
simulation the analytical reduction of the equivalent circuit to its simplest
form is not really necessary, and indeed for computer-aided design it is
5-58
5. E l e c t r ic a l D e s ig n
ptobably better to retain the simple formulation for self- and mutual
indurtance components and to accumulate them into the total phase self
3nd mutual inductances by a suitable programming algorithm.
In the PG-BDC program the accumulation of self- and mutual inductance
components is effected by means of a coilside incidence vector, C. This is
a one-dimensional array in which the number of elements is equal to the
number of slots. Each phase has its own coilside incidence vector. The
element corresponding to the i'th slot is 1 if that slot contains a "go"
conductor belonging to phase 1 , -1 if it contains a "return" conductor,
and 0 if it contains no conductor belonging to phase 1. The total slot-
leak ag e component of the phase self-inductance is then given by

4aOt = ^ N\ L *k i : (5*45)
* Jt-1
in [H], where a is the number of parallel paths and NQis the number of
turn* per coil. It is assumed in equaion (5.45) that the slot permeance
coefficient is different for every slot, allowing for different coilside
positions within the slot. In practice this refinement is rather impractical
and a single permeance coefficient (e.g. equation (5.44)) can be used
outside of the summation.
For the mutual slot inductance between two phases A and B, the
corresponding formula is
= i^ ^ E , W Urua][^CA[^]CB[i-] (5-46)
S jhl
>n [H], where CA and Cg are the coilside incidence vectors for phases A
and B respectively.
5,9,8 End-winding inductance
End-winding inductance is difficult to calculate accurately with simple
formulas because the conformation of the endwindings is complex and
difficult to characterize mathematically in simple terms. Fortunately the
end-winding inductance is generally quite small, and it suffices to have
an approximate formula that includes the effects of the main dimensions.
A simple calculation of end-winding inductance is based on the sum of
the end-winding inductances of each coil, without taking account of the
5-59
5. E l e c t r ic a l D es ic n

Fig. 5.31 Geometry for calculation of end-winding inductance. The same geometry
is used in the calculation of mean length of rum (MLT) and winding
resistance.
mutual coupling between coils. To calculate the end-winding inductance
of one coil, the end-windings are developed into a circular arc as shown
in Fig. 5.31. First, imagine the steel stator core dissolved away. Then,
imagine the coil flattened into a plane, as in Fig. 5.31 b. In c the flat coil
is shown with two semicircular end-windings and two straight sides of
length Z ^ . The diameter of the semicircular end-windings is D = r^a,
and two such semicircles from opposite ends of the motor combine to
make a complete circle whose inductance is
Z,cire *
circ 2 (GMD J
- 2) (5.47)
in [H], where CMD is the geometric mean distance between the
conductors in the coil cross-section. If the coil cross-section is assumed
square, with area A, then CMD = 0.447 -JA. To get the total contribution
from the phase windings to the phase inductance, Lcnd, the inductance
Lcirc is multiplied by the number of coils per phase and divided by a2,
where a is the number of parallel paths.
A more accurate estimate of the end-winding inductance could be
obtained by means of equation (5.17), with the vector potential A

5-60
5. E l e c t r ic a l D esig n
evaluated by a three-dimensional calculation using the finite-clement or
boundary-element method.

5.10 Slotless windings


5.10.1 General
Since the introduction of high-energy magnet materials in the 1970s the
possibility of slotless windings has been of interest The concept is to lay
the siator conductors on the bore of a smooth laminated cylinder, rather
than winding them in slots. Some motors of this type are produced
commercially, and indeed the flat-disk axial-gap motors used in floppy
disk drives are "slotless" also, since the conductors are laid flat on a
prinied-circuit board.
The absence of steel teeth leaves more space available for copper
windings, but it also takes away one of the main conduits for heat
removal and leaves the conductors exposed to the rotating flux. This
raises the possibility of additional eddy-current loss in the conductors,
and of further losses due to circulating currents in any closed loops that
may exist in the winding if there is a varying flux linking them.
The slotless construction permits an increase in rotor diameter within the
same frame size, or alternatively an increase in electric loading without
a corresponding increase in current density. The magnetic flux-density
at the stator winding is inevitably lessened, because of the increase in the
effective airgap length, but the effect is not so drastic as might be
expected. For a motor with an iron stator yoke and an iron rotor body
the magnetic field and its harmonic components can be calculated by the
methods described by Hughes and Miller [4,5]. Considering the
fundamental radial component of B, the value is greatest at the rotor
surface (radius r) and falls off with increasing radius to its smallest value
just inside the stator yoke (radius E), Fig. 5.32. The ratio between the
values of the fundamental space-harmonic radial components of B at these
two radii is given by:
= = ^{rjR fx (54g)
^ i( * ) 1 + (r!R)lp
where p is the number of pole-pairs. If p = 1 (i.e., a two-pole motor) and
t/R = 0.5, this ratio is = 0.4. The ratio for the fcth space harmonic

5-61
D esig n o f bru sh less perm a nent -m a g n et m o t o r s

Fig. 5.32 Geometry of radial-gap ilotlcu motor

can be evaluated from the same formula by putting kp instead of p. For


example, the third harmonic of the two-pole motor is a six-pole
distribution for which fig 0.123. The third-harmonic is attenuated as
the ratio of radii increases. For this example the third harmonic is
reduced to 0.123 of its value at r, while the fundamental is reduced to
only 0.4. The fifth harmonic is reduced to 0.031 of its value at t,
Evidently the rapid attenuation of harmonics gives rise to a more
sinusoidal flux distribution in the stator, which is desirable for sinewave
motors but not for squarewave motors. Of course, the ratio of radii
between the inner and outer layers of the winding may not be very close
to unity, so that the same attentuation causes the outermost layers of thf
winding to experience a flux-density less than that which is experienced
by the innermost layers.
The slotless machine may well accept less expensive grades of lamination
steel because of the absence of slotting and the relatively low flux-density
in the stator yoke. The reactance is also lessened by the reduction of the
slot-leakage inductance, and the risk of demagnetization is decreased by
the decrease in effective coupling between the winding and the magnet.
Cogging torque is eliminated in slotless motors, although there is stilla
possibility for electromagnetic torque ripple when the stator conductors
5-62
5. E l e c t r ic a l D e s ic n
*re carrying current, because these conductors represent discrete
conentradons of current density. However, the attenuation of space-
harmonic flux distributions applies also to the armature reaction flux,
and helps to minimize the torque ripple. The stator conductors may be
skewed by a small amount to eliminate the residual torque ripple.
Because the stator conductors are not constrained to lie parallel to the
axis, it is possible to use a helical winding such as that proposed for
superconducting AC generators [6] or as used in very small PM
commutator motors. Because the helical winding has no end-turns its
utilisation of copper is higher than the severe skew might suggest, and
iE permits the design of a very compact and rugged motor.

Fig. 5,33 Geometry of magnetic field calculation in slotless motor

5.10.2 Design theory for slotless windings


The configuration of a slodess winding is shown in Fig. 5.33. It is
necessary to calculate the flux-density throughout the winding produced
by the magnet, in order to determine the back-EMF. This can be done
using an analytical solution of the field equations, expressed in terms of
5-63
D e sig n o f brush l e ss perm a nent -m a g n et m o t o r s

Fig. 5.34 Representation of permanent magnet by equivalent current sheets at in


edges.

infinite series of space-harmonics. The formulation of this field solution


was developed a long time ago by Bernard Hague, Professor of Electrical
Engineering at Glasgow University [7]. Hagues work was based on the
field of individual conductors and sheets of conductors, so it is directly
applicable to the calculation of winding inductance; but the adaptation
of Hagues technique to work with permanent magnets was published by
Boules [8,9] in 1985. This adaptation is based on the representation oT
a permanent magnet by current sheets along its edges, as shown in Fig
5.34. If the magnet is "hard", i.e. if it has a straight demagnetization
characteristic in the second quadrant, and if it is radially magnetized,
then the sheet current density K is equal to the apparent coercivity fL
of Fig. 4.4. Boules also developed an equivalent current-sheet density
function for arcuate magnets that are parallel-magnetized [8,9]. In this
case there is a sinusoidal distribution of amperes along the inner and
outer surfaces of the magnet also.
Hagues formulas for the magnetic field between two concentric
permeable cylinders were developed in terms of filamentary conductors
rather than current sheets. The simplest way to use Hague's formulas ii
to approximate the current sheet by a small number of current filaments,
as in Fig. 5.33, where each current sheet is replaced by two current
filaments, one at each corner of the magnet. The current at each corner
is 7/2 = KL^j [A], and two corner-currents at the same radius constitute
a "coil". Putting c = (i.e. the surface of the magnet), the radial
5-64
5. E le c t r ic a l D e s ig n
com ponent of flux-density between r = c and r = a, produced by o ne
'coil", is given by
an f c ^ b 2")I r"' s i n c o s nQ (5.49)
cn [a 2j,-b Zn J
where PM is the magnet pole-arc in electrical radians, p is the number of
polc-pairs, and n is the order of each space-harmonic component. If this
u re-written as
fn sin(nPM/p) cos/70 (5-50)
r n-1
then the total radial B component from all "coils" at radius c is
4 = cos[77{0 - ( / - \ )ir lp ]U -\ )n- x (5 -5 1 )
r J- i
A similar calculation is repeated for the 2p "coils" at radius b, and the two
results are added together. The division of each current-sheets into only
(wo comer-currents is coarse, but the approximation is satisfactory
provided that c is not close to a and we are only concerned with the field
in thfi stator winding.
Once the flux-density is known the flux-linkage of any winding can be
calculated. This is a function of the relative orientation between the
magnets and the winding, and therefore equation (5.1) can be used to
determine the back-EMF waveform of the winding as the magnet rotates.
The self inductance of the winding can be calculated by dividing it into
coils, and accumulating the flux-linkage of each coil caused by currents
in each of the coils in turn, including itself. The mutual inductance
between phases can be calculated in a similar way. For this is needed the
magnetic flux-density distribution produced by each coil, and this can be
determined using the equation corresponding to equation (5.49), but for
r< c instead of r> c. Instead of the magnet equivalent corner-current
7/2, the actual coil current is substituted. This equation is given by
Hague in Ref. [7]. End-winding contributions can be added, as in
section 5.8.6.
The calculation of the field, back-EMF waveform, and winding
inductances for completely arbitrary distributions of conductors,
including helical windings and even Gramme-ring windings, can be
5-65
D esig n o f brushless perm a nent -m a g n et m o to r s

undertaken using three-dimensional finite-element analysis and boundary-


element analysis programs, which accumulate the winding flux-linkage
using equation (5.17) after solving for the vector potential.
References
1 Kenjo T and Nagamori [1985] Prrmanemt-magnet and brushless DC molars, Oxford
University Press.
2 Alger PL [1970] Induction machines, Gordon and Breach, NY
3 Veinott CG [1959] Theory and design of small induction motors, McGraw-Hill, NY
4 Hughes A and Miller TJE [1977] Analysis offolds and inductances in air-corni and
iron-corrd synchronous machines, IEE Proceedings 124, 121-8
5 Miller TJE and Hughes A [1977] Comparative design and performance analysis of air-
corrd and mn-corrd synchronous machines, IEE Proceedings 124, 127-32
6 Ross JSH [1971] UK Patent 1395122
7 Hague B [1962] The principles of electromagnetism applied to electrical machines -
republication by Dover Publications Inc. of Electromagnetic Problems in Electrical
Engineering [1929]
8 Boules N [1985] Prediction of no-load flux density distribution in permanent magnet
machines, IEEE Transactions, Vol. IA-21, No. 4, May/June 1985, pp. 633-643.
9 Boules N [1984] Two-dimensional analysis ofcylindrical machines with permanent magnet
excitation, IEEE Transactions, Vol. IA-20, No, 5, September/October 1984, pp. 1267-
1277.
10. TJE Miller [1989] Brushless permanent-magnet and reluctance motor drives, Oxford
University Press, ISBN 0-19-859369-4

5-66
6. SINEWAVE MOTORS
6.1 Introduction

6.1.1 The ideal sinewave motor


The squarewave motors considered in Chapter 5 had the following ideal
chacterisdcs:
1. rectangular distribution of magnet flux in the airgap;
2. rectangular current waveforms; and
3. concentrated stator windings.
The sinewave motor differs in all three respects. In its ideal form it has:
1. sinusoidal distribution of magnet flux in the airgap;
2. sinusoidal current waveforms; and
3. sinusoidal distribution of stator conductors.
We begin with a study of the properties of the ideal sinewave motor,
developed initially from the basis of perfectly sine-distributed windings.
This differs from the classical theory of AC machines, in which the
fundamental winding element is the single concentrated full-pitch coil.
In classical theory, the all-important fundamental flux, and the torque
and EMF associated with it, can be extracted only after the harmonic
winding factors are understood and applied, whereas in the present
approach these essential parameters emerge without complication from
simple integrals. The harmonic winding factors are developed afterwards,
in order to characterize practical windings in terms of the ideal sine-
distributed windings.
A more important reason for beginning with the ideal sine-distributed
winding is that it is closely and simply related to the concept of the space
vector, which is probably the most compact form for expressing the
performance equations of AC machines. Space vectors of flux, flux-
linkage, current, MMF, EMF, and voltage are single complex numbers
6-1
D e sig n o f bru sh less perm a n en t -m a g n et m o t o r s

that represent the magnitude and direction of the the spatial distribution
of these quantities in a polyphase machine with a rotating field. They are
analogous to the time phasors which represent the magnitude and phase
of individual phase currents, voltages, etc., and they are closely related
to them. The space vector is also closely related to the d,q-axis frame of
reference, which is the basis of classical AC machine theory. Space
vectors actually have a long history, but they are increasingly important
and popular today because of the widespread use of vector control or fieId-
oriented control for synchronous and induction motors. Certain pulse-width
modulation (PWM) control strategies used with vector control are more
easily understood and analysed in terms of space vectors, than in terms
of tfj^-axis theory.
6.1.2 Practical motors designed to approximate the sinewave motor
The most fundamental aspect of the sinewave motor is that the back-EMF
generated in each phase winding by the rotation of the magnet should
be a sinewave function of rotor angle. The purity of the sinewave
depends partly on the magnet flux-distribution, which should be as near
as possible to a sinewave, and pardy on the winding distribution. If the
winding were perfecdy sine-distributed it would have no flux-linkage with
space-harmonics of the magnet flux-distribution, but practical windings
are not perfecdy sine-distributed and therefore it is important to make
the magnet flux distribution as nearly sinusoidal as possible. The rotor
configurations shown in Fig. 6.1 are commonly employed for this.

(a ) (b ) (c )
Fig. fi.l PM rotors commonly used in sinewave motors.
In Fig. 6.1a the magnets are parallel-magnetized (not radially
magnetized). In Fig. 6.1A the varying thickness of the magnet in ihf
direction of magnetization naturally profiles the flux-distribution and a
6-2
6. S inewave M o t o r s
very good sinewave is possible with this configuration. In Fig. 6.1 c the soft
iron pole-pieces can be profiled to give a variable airgap length which
produces the same effect. A difficulty with this motor is the effect of
cross-magnetization, i.e. 9-axis armature reaction flux. In all three cases
Ihc magnet pole arc is chosen to maximize the ratio of the fundamental
flux to the total flux.
The windings can be made approximately sine-distributed by three main
methods:
1. short-pitching or "chording";
2. skew; and
3. distribution or "spread".
Short-pitching means winding the coils with a span less than n electrical
radians; this has the additional advantage of lowering the resistance and
decreasing the amount of copper in the end-windings, as well as making
the end-windings more manageable in the factory. Winding pitches of
5/6, 2/3, and even 1/2 are typical, the fractional pitch being relative to
one pole-pitch or 7t electrical radians. "Concentric" windings are
essentially made up of combinations of short-pitched coils, all of which
have the same axis.
Skew can be applied to either the winding or the magnets, and both
methods are used in production.
The distribution or spread of a winding means that the conductors are
distributed throughout an angular belt as discussed in connection with
Fig. 5.13. In large AC machines the spread is achieved by means of a lap
winding, in which all the coils are identical. In small PM machines it is
more usual to use concentric windings, of the type discussed in Chapter
5.
The sinewave PM motor is a simple synchronous motor. It has a rotating
stator MMF wave and can therefore be analyzed with a phasor diagram;
this is especially useful in designing the control system and calculating
the performance. In this chapter we determine expressions for the
torque; the open-circuit phase EMF due to the magnet; the actual
winding inductance; and the synchronous reactance. These results are
6-3
D e sig n o f b ru sh less perm a nent -m a gnet m o t o r s
modified for practical windings by means of the standard winding factors
of AC machines. They provide the basis for the phasor diagram, which
is used to develop the circle diagram and study its variation with speed;
from this the speed/torque characteristic is derived. It is shown that the
surface-magnet sinewave motor has limited capability to operate along a
constant-power locus at high speed. The chapter finishes with a
theoretical comparison of the torque per ampere and kVA requirements
of squarewave and sinewave motors and a comparison of wound-field and
PM motors which justifies the preference for PM motors in smaller sizes.

6.2 Properties of sine-distributed windings


6.2.1 Conductor and Ampere-conductor distributions
The ampere-conductor distribution of a sine-distributed winding is
expressed by the sheet current density
K = i sin/70 (6-D
in ampere-conductors per radian or plain amperes per radian, The
number of conductors per pole is

1 = ^ (6.2)
2 v p p
so the total number of conductors per phase is Ns/p x 2p = 2Nt. The
total number of turns per phase is half the number of conductors per
phase, i.e., Ns. This is illustrated for a two-pole winding in Fig. 6.2.
6.2.2 Airgap flux produced by sine-distributed winding
For the calculation of back-EMF and winding inductance it is necessary
to calculate the flux-linkage of the sine-distributed winding. Fig. 6.2 shows
the self-flux produced by current in the winding itself, and Fig. 6.3 shows
the MMF which forcesthe flux across the airgap at two points located at
0 and n /2 p - 0 electrical radians. The equations will be developed for
the general case of a machine with 2p poles, even though Fig. 6.2 b
drawn for p = 1. The MMF enclosed by the flux line is given by

6-4
6. S inew ave M o t o r s

FLUX LINES

COIL A X IS

ELEM ENTAL
COIL

]% fi.2 Two-pole sine-distributed winding; conductor and ampere-conductor distribution.

wip-e v . iN . (6.3)
F = f i smp6 d9 = -----cospO.
Je 2 p
The airgap flux-density at 0 is therefore
B = MoH ~ = B cospO

b = (6.5)
2p g "
and g is given by equation (5.17). The flux linking the coil whose
conductors are located at the angles 0 and -0 is given by

<f> = jr* B* cos p9 d6 = $ sin pO, (6.6)

6-5
D e sig n o f bru sh less perm a nent -m a g n et m o t o r s

Fig. 6.3 Two-pole sine-distributed winding: calculation of MMF.

where $ is the flux per pole and is given by


$ = P DL (6.7)
P
in [Wb], with D = 2^; this is obtained from equation (6.6) with 0 =
n / 2p, corresponding to the full span of the coil (it electrical radians).
This is only 2/ n times the flux that would be obtained if all the ampere
conductors were concentrated into a full-pitch coil; (see equation (5.22)).
6.2.3 Self flux-linkage and inductance of sine-distributed winding
The fluxlinkage of an elemental coil whose conductors are located at
the angles 0 and -0 (Fig. 6.4) is given by the flux linked, multiplied by
the number of turns in the elemental coil, i.e.,
jy
dty = $ sin p8 * ~ sin p6 d9, (6-8)

6-6
6. S inewave M o t o r s

Fig. 6.4 Two-pole sine-distributed winding: calculation of flux and flux-linkage.

so the total flux-linkage of the winding, with all 2p poles in series, is


(6.9)
Perfect linkage would be jV4>, so with sine-distribution the winding links
only 78.5% of its own flux. The inductance is now easily derived as i|i/i,
and after back-substituting for $ , B and F we get
7T PoJ^ 2^'stk/l (6 .10)

in [H]. If there are a parallel paths through the winding, this formula
ii unchanged provided that Af is interpreted correcdy as Nc x 2p/a,
where Nc is the number of sine-distributed turns per pole. Ns is then the
number of turns in series per phase, in the same sense as in Chapter 5,
section 5.8.3. Comparing equation (6.10) with equation (5.23), the sine
distribution reduces the self-inductance by 50% compared with that of
a single concentrated full-pitch coil having the same number of
conductors. As noted already, the flux is reduced by the factor 2/ n while
the linkage of that flux is reduced by the factor ir/4, so that for the same
ampere-conductors the inductance is reduced by 2/ti x ti/4 = 1/2,
compared with the concentrated coil.

6-7
D esig n o f b ru sh less perm a n en t -m a g n et m o t o r s

Fig. 6.5 Mutual inductance between sine-distributed windings

6.2.4 Mutual inductance between sine-distributed windings


Suppose the flux is established by another sine-distributed winding [J, of
the same diameter, but having a different number of pole-pairs q, and
suppose that the axis of the exciting coil is displaced y electrical radians
from the axis of the first coil, which is designated a, Fig. 6.5. Then by
generalizing the calculation in equations (6.8) and (6.9) we can calculate
the flux-linkage of winding P due to current i in a as
(6.11)

and this integral is zero unless p = q, in which case


Y/j. = >s Y. ( 6 . 12)

From this it is easy to show that the mutual inductance between two
otherwise identical sine-distributed windings with Na and JVp series turns
per phase is
cos y = Mjnm cos Y. (6-13)

6-8
6. S inew ave M o t o r s
where Y is the angle between the axes of the windings, in electrical
radians. These results are among the most fundamental in the theory of
AC machines. The vanishing of the integral in equation (6.11) for p *
q means that a sine-distributed winding has no flux-linkage with any sine-
distributed flux that has a pole-number different from its own: by
implication, it therefore rejects all space-harmonics of the magnetic flux
set up by the rotor magnet, since this field is no different in principle
from the field set up by another winding or set of windings. The sine-
distributed winding is therefore a perfect "notch" filter for space-
harmonic fluxes of its own pole-number.
The cosinusoidal variation of the mutual inductance in equation (6.13)
is equally fundamental. In particular, it forms the basis of the two-axis
theory of the induction machine, and it also plays a central part in the
two-axis (dj^-axis) theory of synchronous machines, including the PM
machines of this chapter. If y = 0 and Nat = Np = Ns, then the mutual
inductance Map is equal to the self-inductance of equation (6.10).
6.2.5 Generated EMF
Suppose the flux is established by a second winding which rotates relative
lothe first winding. The flux-linkage in the stationary winding ais still
given by equation (6.12) but with

Y = t V + Yo (6-14>

where oij is the angular velocity in electrical radians/sec. The EMF


generated in the a winding is given by Faradays Law:
e = = * ( *+ Yo)- (615)
The coefficient of the sine term is the peak EMF. If Na is replaced by
A; and is replaced by <I>j, the general expression for the RMS value
t)f the fundamental generated EMF is obtained:

E ' a ( ? " ' ) ( <U6)


in [V]. Note that is the fundamental flux per pole, not the total flux
per pole. The main flux is usually established by the magnet, rather than
by a sine-distributed winding.

6-9
D esig n o f bru sh less perm a nent -m a gnet m o t o r s

6.2.6 Torque
Suppose the flux is established by a second winding or a magnet with y
pole-pairs, whose axis is located at an angle p from the reference axis of
the first winding, Fig. 6.5. The radial flux-density at the winding is
B cos ( g9 - f3) and this produces a force on the winding element and
a resulting torque equal to
= y, * B cos (g$ - /9) * N sin pO * (6.17)
The total torque is the integral

e 2 (6-18)
^ f {sin [{p + g)6 0] + sin [ (p - g)8 + 0]]dB.
22 Jn

This integral is zero unless p = q, in which case


BL^D
---Ji T_ iN.
71 = ---- x - 2 tt sin /}
e 4 2
= A * 3>, * p * i sin
V Kr j,. pa (6.19)'

ty jp s m fi
-

in [Nm]. This equation describes torque produced by the interaction of


the sinusoidal distribution of ampere-conductors and the fundamental
flux produced by the magnet (or by another winding). A pure sine-
distiibuted winding acts as a perfect notch filter in that it produces no
average torque with space-harmonics of the flux.
6.2.7 Rotating flux and ampere-conductors
If pQ is replaced by (pQ - then both the flux and the ampere
conductor distribution rotate at synchronous speed s electrical
radians/sec. The rotation of the flux is achieved by physical rotation of
the rotor, so that
B(9) = cos (p O -p ). <6'20)

6-10
6. S inew ave M o t o r s
The rotation of the amperc-conductor distribution is achieved by means
of a two- or three-phase winding carrying balanced currents. If three
phase windings a,b,c have their axes at 0, + 120, and - 120 and are
supplied with currents i cos 0)at, cos (&)at - 2tc/3), and i cos (Ci>st -
2n/3), the resulting ampere-conductor distribution is given by
N [sin p 9 costds/ + sin (p Q - 2tt/3) cos {(i3s/ - 2 tt/ 3 )
+ sin (p6+ 2ir/3) cos (ws/+2ir/3)] (6.21 )
t N
= / sin (p9 -
2 2 s

In normal operation the stator supply frequency (in rad/sec) is made


equal to the rotation frequency, i.e.,
ws = 2ir/ (6.22)
in electrical rad/sec, where / is the supply frequency in Hz. The
mechanical angular velocity is
om = (6.23)
in [rad/sec]. With three phases and balanced sinusoidal phase currents
the torque is given by equation (6.53) below.
6.2.8 Vector control or "field-oriented, control"
The stator ampere-conductor distribution rotates in synchronism with
the rotor and the torque angle is kept constant, usually by means of a
simple form of vector control or field-oriented control which requires
a shaft position sensor (i.e., encoder or resolver feedback). If the supply
frequency and the rotation frequency were unequal, the motor would be
running asynchronously. No average torque would be produced, but
there would be a large alternating torque at the beat frequency or
pole-slipping frequency.
In vector control, v|/ and i are controlled independently: indeed in many
cases 1|> can be considered fixed as it is often set up by a high-coercivity

6-11
D e sig n o f bru sh less perm anent -m a gnet m o t o r s

magnet. The torque angle p is kept at 90 in order to maximise the torque


per ampere, and this is done by phasing the current waveforms relative
to the rotor position. The term "orientation" refers to the direction of
the spacc-vector of current, which is described later. With this type of
control the AC permanent-magnet brushless motor obeys the relationship
Torque * Flux * Current (6.24)

This applies with AC current in the same way as it applies with DC


current in the DC motor and in the brushless squarewave motor. In the
AC motor the flux is constant, sinusoidally distributed in space, and
generates a sinusoidally varying EMF in each phase. An observer on the
rotor, travelling at synchronous speed, would "see" perfectly constant flux
and current when the motor is operating in the steady state.
6.2.9 Synchronous reactance
The flux and flux-density in the torque equation are those produced by
the magnet acting alone; in other words, they are open-circuit values and
do not include any contribution due to the MMF of the stator currents.
Although the armature-reaction MMF modifies the airgap flux-density, iL
does not figure in the torque expression unless it significantly affects the
saturation level of the magnetic circuit. Physically the stator may be
regarded as being incapable of producing torque on itself. The armature
reaction flux is aligned with the stator ampere-conductor distribution and
therefore has an effective torque angle of zero. The armature reaction
flux does, however, rotate and it induces a voltage drop in the phase
windings, which must be overcome by the supply voltage; this ii
accounted for by the synchronous reactance.
The rotating ampcre-conductor distribution of equation (6.21) sets up
a rotating flux wave
a cos (p9 - u>st) (6-25)

where

(6.26)

and I - /V2 is the RMS phase current. The associated flux per p o le is

6-12
6. S inew ave M o t o r s
/A rotating flux wave, established by armature reaction,
generates voltages in all three phases. In each phase the voltage is
proportional to /an d is therefore regarded as the voltage drop ^/across
a fictitious reactance, the 'synchronous reactance, By substituting the
flux/pole into the expression derived earlier for EMF, and dividing by I,
we get
y _ 3 tt Mo^i-^ak (6 27)
^ 8 p ig "
in [Ohms], This expression applies to an ideal 2/>-polc sine-distributed
three-phase winding with Af turns in series per phase, and it neglects the
leakage inductance of the slots and end-turns. To obtain a practical
formula for a real winding, we must first find an effective value for the
iinc-distributed turns. This is done by means of Fourier analysis and
winding factors in the next section.

6.3 Real windings


The most important characteristic of AC windings is how much
fundamental flux they produce; or, if they are generating EMF from a
rotating flux wave, how much fundamental flux they link or "extract"
from the flux wave.
6.31 FuU-pitch coil
Fig. 6.6 shows what is effectively a single full-pitch coil in a 2-pole
marhine. In order to preserve symmetry and to ensure that the 2-pole
winding is treated in the same way as windings with higher pole numbers,
the coil is divided into two equal halves, one associated with each of the
rvro poles. The number of turns per pole is Np, and the number of turns
in each coil is NQ. Since there is only one coil per pole, = Nc. The
number of conductors per pole is 2A' (= 2Nc). The distincUon between
At, and Nc may seem unnecessary in this case, but they become unequal
wnen the number of coils per pole is greater than 1 .
The ampere-conductor distribution in Fig. 6.6 generates a rectangular
flux distribution in the airgap with a peak (flat-top) value B as in Fig.
5-22. The MMF distribution is also rectangular with peak value N^i By
Fourier analysis the fundamental component of this distribution is =
6-13
D esig n o f bru sh less perm a nent -m a gnet m o t o r s

(4/n)ATpi and this implies that the effective number of sine-distributed


turns/pole is (4/n)iVp. In other words, a sine-distributed winding
carrying the same current would need (4/it)iVp turns/pole in order to
set up the same fundamental flux. If the number of pole-pairs is p, and
all turns are in series, the total number of turns/phase is 2pNp. If these
turns are reconnected in a parallel paths, the number of turns in series
per phase is ATph = 2pNp/a. In the equivalent sine-distributed winding the
number of scries turns per phase would be Ns = (4/n)N ph.

Fig. fi.6 Full-pitch coil, shown divided into two equal halves, one per pole

6.3.2 Short-pitch coil


Fig. 6.7 shows a pair of short-pitch coils (also known as chorded roils),
with one coil per pole. The flux distribution is shown in Fig. 6.8. By
Fourier analysis the fundamental component of this distribution is given
by
= sin% B. (6.28)
31 7T 2

6-14
6. S inewave M o t o r s

2N p C O N D U C T O R S /P O L E

N T U R N S /P O L E
P

N c TU R N S /C O IL

Tig 6.7 Two short-pitched coils in 2-pole machine

FUNDAMENTAL

SPAN = 7 SLOTS

Fig 6.8 Flux distribution produced by short-pitch coils of Fig. fi.7


where
B = (6.29)
8 g"
and Np = Nc conductors/pole.
6-15
D e sig n o f brushless perm a nent -m agnet m o t o r s

The fundamental pitch factor is now defined as the ratio


Bx produced by short-pitch coil or
T B x produced by full-pitch coil ^
Aii
"pi - sin 2 -
. OL
cos.2
The angle 0 is the pitch or span of the coil, and e is the chording angle.
both these angles are in electrical degrees or radians. For the winding
shown in Fig. 6.7, a = 7 ti/ 9 , i.e. the coil span is 7 /9 , and e = n / 9 . The
fundamental pitch factor is sin (1/2 x 7 tc/ 9 ) = 0.93969, which means
that the winding will produce only about 94% of the fundamental flux
of a full-pitch winding with the same number of ampere-conductors.
Short-pitching is an important means for eliminating harmonic linkages.
The n tk harmonic pitch factor is defined as
ZjJ, produced by short-pitch coil
-*pn B^ produced by full-pitch coil (0.31)
^pn sin n2 = cos n 2~ .
a e

Bn is the nth space harmonic of the flux distribution. The nth harmonic
can be eliminated from the flux wave produced by a winding if = 0,
which requires e = n /n . In the winding of Fig. 6.7, 0 = 2n/9, so no
integral-order harmonic is eliminated- If the coil span was reduced from
7 to 6 slots, however, e would be 3n/9 = n/S and the third harmonic
would be eliminated from the self-flux because = cos (3 x ti / 6) =0.
The winding would also have zero linkage with the 3rd harmonic flux
produced by any other winding or by the magnet. Short-pitching by 1/n
of a pole-pitch eliminates the nth harmonic flux-linkage from all sources,
both self and mutual.
In wye-connected three-phase motors there is little point in using a
chording angle e that is n /3 (or a submultiple thereof), because the wyr
connection itself cancels the 3rd harmonic EMFs from the line-line EMF,
and no third-harmonic current can flow provided that the star point of
the winding is isolated. However, in delta-connected motors it is essential

6-16
6. S inew ave M o t o r s
(0 eliminate the third harmonic (see section 5.2.2), and the use of a 2/3-
pitch winding does this most effectively.
,\ concentric winding with more than one coil per pole contains coils of
differing span, and even different numbers of turns. Provided that all the
coils have the same winding axis, the overall pitch factor can be
calculated using
_ -^i^pn(l) + pn(2) + + -^m^pnfrn) /fi 391
^ ----------------- a w v T a T------------ (6-32)
where m is the number of coils, JVj, M,, etc. are the numbers of
turns/coil, and i n(1j, fcpn(2)> etc- are t^e individual pitch factors for the
individual coils. For example, a winding of 2 coils per pole having spans
of 7/9 and 5/9, with equal numbers of turns, would have a third-
harmonic winding factor of
sin ~3 -----
7 TT + sin -3 x -----
5 7T
92 92
2 (6.33)
sin (210*) + sin (150;)
2
= 0.
This shows the possibility of eliminating a particular harmonic by means
of a "composite" pitch factor, even though no coil has the exact pitch
required to eliminate that harmonic from its own flux-linkage.
6.3.3 Distribution or spread
Lap windings are made up of groups of coils which all have the same
span, but which are displaced from each other by an electrical angle y
(see Fig. 5.5). Often y is equal to the slot-pitch angle, but this is not
necessarily the case, particularly in fractional-slot windings.
Consider two coils such as those in Fig. 5.5. Each produces a
fundamental airgap MMF that is sinusoidal and whose wavelength is, by
definition, 2n electrical radians. The two fundamental sinewaves are
displaced from each other by y electrical radians, so that the resultant
airgap MMF is proportional to
sin + sin (0 - Y) = 2 sin cos (6.34)

6-17
D esig n o f b ru sh less perm a n en t -m a g n et m o t o r s

If both coils were concentrated together in the same slots then y = 0 and
the MMF would be proportional to 2 sin 0. The ratio of the amplitude
of the fundamental of the distributed winding to that of the concentrated
winding is cos ( y / 2), and this is known as the fundamental distribution
factor, fedI. If there are m coils in a group, the resultant fundamental
MMF is proportional to
sin 8 + sin(0 - y) + sin(0 - 2 y) +...+ sin [9 - (J37- 1) Y]
s in ^ l
= sin 8 + (in~Y) v2
(6.35)

2 sin-^-
2
(This result can be derived by multiplying both sides of the original sum
by 2 sin (y / 2) and expanding the sin-product terms into sum- and
difference sin terms, then cancelling common terms). If all the m coils
were concentrated together then y = 0 and the MMF would be
proportional to m sin 0. The fundamental distribution factor is then
given by the ratio
sin m y
I
m sin Y
[2 J
The same procedure works also for the space-harmonic components of
the MMF, so that the nth harmonic distribution factor is
sin 'nBH.
2 (6.37)
m sin n l
2
Equation (6.35) shows that the axis of the resultant MMF wave ij
displaced by an angle (m - 1 ) y /2 by the distribution, relative to ihc
position it would have if all the coils were concentrated together with the
first coil. The summation in equation (6.35) can be represented
vectorially as in Fig. 6.9, where the vector sum of the MMFs is
represented by the large chord.

6-18
6. S inew ave M o t o r s

fi.9 Vector sum of coil MMF'j, illustrating the winding distribution factor

If all the m coils were concentrated their resultant MMF would be


represented by m times one of the small chords. The ratio of the large
chord to the product of m times one of the small chords is given by
equation (6.36). Fig. 6.9 is often used to derive or illustrate the
derivation of the fundamental distribution factor. Usually the diagram is
used in relation to the sum of the EMFs generated in the coils by the
passage of a sine-distributed flux. The distribution factor and indeed all
the winding factors are the same for the MMF and the EMF. If the coils
are laid in adjacent slots the angle my is called the phase spread.
63 4 General case
The pitch and distribution factors described above can be readily applied
to regular winding patterns such as concentric or lap windings. However,
it is possible to wind "custom" windings which do not conform to these
standardsparticularly fractional-slot windings. For these cases it is
desirable to be able to calculate a general winding factor j for the
fundamental and Apd n for the harmonics, for any distribution of coils,
including coils of differing pitch. Referring to the single coil in Fig. 5.25,
the MMF-distribution ^ 0 ) is similar to the one shown for coil Gl-Rl in
Fig. 5.27, and the values of the flux-density in the "infield" and in the
"outfield" are given by equation (5.36) with corresponding values of the
MMF, Fj and FQ. The full Fourier-series representation of the MMF
distribution i^Q) is given by
00
F{8) = [^cos n6 + Z^sin n8] (6.38)

6-19
D esig n o f bru sh less perm a nent -m a gnet m o t o r s

where
sl = f F{6) cos n& d6 and
2 ir Jo
(6.39)
IT / n 7^0) sin nO d9.
2
O.ir j

If wc add the subscript q to denote the <y'th coil of a phase winding,


remembering that "G" stands for the "go" conductor and "R" stands for
the "return" conductor, then
*n.q = (-^ofl)cos d

+ J (/j^cos jj0
fl<M (6.40)
r27r
* J (~Foq)cos 1,9 d6

2
(Nf)a cos n<pnnai
sin
nv 4
and
(Nf)a sin n<f> sin (6.41)
IITT M 2
where a = (9Rq ~ 0G ) is the span of the coil in electrical radians, (|>^
= ( 0R + dGc[)/2 is the location of its axis, and (M)q is the ampere-turns
of me i/th coil. For a winding with N coils the overall harmonic
coefficients are given by
E a n,q At = E Awi (6 '42)
q-l ^1
and the magnitude of the nth harmonic is
^ = W * V- (6-43)
If all the ampere-conductors were redistributed and concentrated in full-
pitch coils, one per pole, then the MMF distribution would be a
rectangular wave of peak value F /2, where

6-20
6. S inewave M o t o r s
n
(6 .4 4 )

The amplitude of the nth harmonic component is 4/y'wi, and this is the
base from which the nth harmonic winding factor can be defined:

The evaluation of an, bn, cn and thence fcpd n is straightforward in


computer-programming terms, if the coil table contains details of the
number of turns and the "go" and "return" slots occupied by every coil.
These equations are valid if all angles are measured in electrical degrees
or radians.
6,3.5 Skew
Either the winding or the magnet may be skewed. Skew is usually used
to eliminate cogging and tooth-ripple effects which cannot be eliminated
by winding distribution. (See Fig. 8.3 and Fig. 8.4). Indeed with an
integral number of slots per pole per phase the pitch factor for the
tooth-harmonics is equal to that for the fundamental. The n'th harmonic
Aon factor is defined in the same way as the pitch factor, i.e.,
produced by skewed winding
Bn produced by unskewed winding
(6.46)
*sn

In the machine of Fig. 6.7 with 18 slots, a skew of 1 slot-pitch gives Asl =
0.995, = 0.0583, and fcjig = -0.05326, and these low skew factors
attenuate the most troublesome harmonics associated with slotting. A
skew of one slot pitch is not necessarily optimum from all points of view,
and smaller values are sometimes used, e.g. 0.75 of a slot-pitch.
6.3.6 Design formulas for practical windings
All the design formulas for sine-distributed windings can now be
rewritten for practical windings simply by replacing the sine-distributed
6-21
D e sig n o f bru sh less perm anent -m a g n et m o t o r s

series turns per phase with the equivalent sine-distributed series turns
per phase for the fundamental:

7T

where Nph is the actual number of series turns per phase 2pNp/a, and
is the fundamental winding factor.
K i = *pdi^i- (a48)
The open-circuit phase EMF is
E - A (6.49)
J2
where $>M1 is the fundamental flux due to the magnet. If, as is usually the
case, the magnet flux distribution is not perfectly sinusoidal, the
fundamental component should be extracted using Fourier analysis as
outlined in section 6.4 below. In that case the nth harmonic EMF can
be computed using equation (6.49) with Awl replaced by kmi. The EMF
can be written in an alternative useful form by recognizing that the RMS
phase flux-linkage due to the magnet is
^iul
Wl ph Ml (65Q )
Ml =
v/2
With this, the EMF can be written
Eq = wsl).M1. (6.51)
The subscript 'q here means that the EMF phasor is aligned with the
<f-axis in the phasor diagram since it leads the flux-linkage by 90 electrical
degrees.
The torque is given by equations (6.19), (6.21-23) and (6.50):
T, = 3 ^ mi I P sin P (6.52)

6-22
6. S inew ave M o t o r s

E J s in /3 (6.53)
q
jn [Nm). o>m is the mechanical angular velocity given by equations (6.23)
and (6.22). For a 2-phase motor the 3 in equations (6.41) and (6.42)
should be replaced by 2.
Turning now to the actual phase inductance, it is meaningless to
substitute the effective sine-distributed turns. The actual inductance must
include all the self flux-linkage, not just the fundamental component.
The inductance formula for a practical winding therefore remains the
same as in section 5.8, including the leakage inductance.
The synchronous inductance or reactance is quite different from the
actual inductance, being associated with the voltage drop in one phase
caused by the fundamental component of rotating armature-reaction
flux, under balanced conditions with all three phases in operation.
Substituting for the effective sine-distributed turns per phase, we get
(6.54)
in Ohms per phase. To this value must be added the per-phase leakage
reactance Xa, giving the total synchronous reactance
(6.55)

6.4 Salient-pole motors


6.4.1 Calculation of X^
PM rotors with salimcy have / X^. To understand this physically,
consider the armature-reaction flux from the point of view of an observer
on the rotor. The ampere-conductor distribution and the armature
reaction flux both rotate at synchronous speed and therefore appear
stationary to an observer on the rotor.
The rotor has two axes of symmetry, the pole-centre axis or d-axis and
the interpolar axis or ^axis. The effect of the armature ampere
conductor distribution and MMF in setting up armature-reaction flux can
6-23
D esig n o f b ru sh less perm a nent -m a g n et m o t o r s
be analysed by considering the MMF to be resolved into a rf-axis
component Flad cos and a 9-axis component FlatJ sin pQ. These MMF
components are the fundamental components of the actual ampere
conductor distribution. The dr and 7-axis fluxes which they set up are not
sine-distributed because the airgap permeance is not uniform. Moreover,
in the d-axis the airgap permeance is in series with the magnet
permeance, which complicates the magnetic circuit somewhat.
For the calculation of synchronous reactance it is the fundamental
component of the resultant armature-reaction flux that is required. This
can be obtained from the actual flux by Fourier analysis, as shown in this
and the next section. The actual or total fluxes in the d- and faxes are
also important from the designers point of view, for separate reasons,
and these are discussed in sections 6.4.3 and 6.4.4.

Potential
^ F ladc o sP e

1 magnet
Potential u f 1------ n
Zero potential 2p
Fig. 6.10 Geometry for the calculation of XA.

The geometry for the calculation of is shown in Fig. 6.10 and the
axis flux distribution in Fig. 6.11.

6-24
6. S inew ave M o t o r s

Fif 6.11 liaxii airgap flux distribution

This geometry is derived from the spoke-type motor shown in Fig. 6.1 (c),
by developing it into Cartesian or rectangular form. The projection is
used purely to facilitate a clear picture and does not introduce any
additional approximation, since the circular shape of the airgap is
retained in the mathematics. Although the geometry appears to be
specialized to the spoke-type rotor, in fact it applies to all rotor
c o n f i g u r a t i o n s provided that the symmetries of the magnetic circuit are
correcdy observed. Fig. 6.12 shows how the basic geometry of Fig. 6.10
can be developed for three different rotor configurations. Similar
developments apply to the geometry for calculating X^.
The MMF distribution fjad cos pB is determined from Fig. 6.13. The
actual winding has 2p poles, and each pole has N turns. For example,
if there are m coils per pole each with Nc turns/coil, = mNc. The
equivalent sine-distributed winding has a conductor distribution (Ns/2 )
tin pQ conductors/rad, so that the number of conductors per pole is
v = fJirfp _Vs sin p6d6 = i 2x - Ai n lw = N-, if-
t6-56)
v Jo 2 ir 2 p p
6-25
D esig n o f bru sh less perm anent -m a gnet m o t o r s

M A G N ET

Fig. 6.12 Development of geometry for calculation of X ^. (a) Spoke type (b) Erat
type (c) Surface magnet
6-26
6. S inewave M o t o r s

fig. 6,13 Actual and equivalent line-distributed windings

and the number of sine-distributed turns per pole is half this, i.e.

N = (6.57)
p 2 Ip
If all turns are in series then the number of sine-distributed turns in
series per phase is
2p * % = Ns. (6.58)
Ip s
In a practical phase winding with all Arph turns in series,

N. - +7r K , <6-59>

From Fig. 6.13 the MMF per gap has a peak value

6-27
D e sig n o f brushless perm anent -m a gnet m o t o r s

K
F* ' N* t P

2 2 p
(6'M)
3y/2 AvI
IT
ampere-conductors/gap. This equation shows that in a multiple-pole
machine the total armature-reaction ampere-tums are divided among p
pole-pairs. If the phase winding is divided electrically into a parallel
paths, then the MMF per gap is divided by a, and this is taken into
account by defining JVph as the number of turns in series per phase, i.e.
N _ T o ta l turns/phase
Pb " Paths

Poles x Coils/pole * tunis/coil (6 61)


Paths
2p m Nc

The peak MMF per gap given by equation (6.60) is made equal to
^lad by using the rf-axis current 7d from the phasor diagram (section
6.5), or to by using the y-axis current 7^.
Returning to the magnetic analysis of Fig. 6.10, the pole-piece is floating
magnetically and has an undetermined magnetic potential u. Since the
q-axis is taken as the datum of magnetic potential, is themagnetic
potential drop across one magnet. "One magnet" isdefined as the
segment of magnet associated with one half-pole. In the spoke-type rotor
this means splitting the actual magnets into two sections in series. The
permeance P ^ is the permeance of "one magnet", i.e.
Am/ (L ra/ 2 ) , where Lm is the total thickness of one block of magnet. The
airgap flux-density over the pole-piece is given by
3 .dW = [^iad COS P - (6 62)
g
6-28
6. S inewave M o t o r s
and the form of this distribution is shown in Fig. 6.11. If the pole-piece
is sufficiendy wide and the negative potential-drop u induced in the
airgap by the reluctance through the magnet is sufficiendy large, a
proportion of the flux can re-emerge from the pole-piece and return
across the airgap in the reverse direction, completing its path to the q-
axis via the stator teeth at the edges of the pole arc. These re-entrant
flux-paths were termed "whorls" by V.B. Honsinger in his analysis of line-
siart versions of the PM AC motor [1],

The total d-axis flux corresponding to ad(0) is


a ir flp
*ad = 2 J B ^m rL ^dO
0 (6.63)
= 4 - ^ a d ^ d - ]
where
. = sin a n /2 (6 64)
air!2
and
R, iL ___ = 2p*'
8 ll a n D L \^ a n D L ^ (6-65>
^ 2p *
and D= 2 ris the stator bore diameter. R^ is the airgap reluctance taken
over the whole arc of the pole-piece.
The flux $ ad is equally divided between the two halves of the pole, and
*ad/2 goes through the parallel combination Pm0][PL (FiS- 4-7)> where
is the permeance of one magnet as shown in Fig. 6.10 and PL is the
leakage permeance associated with that magnet. Both ^m0 and PL are
confined to one half-pole, defined by the axis of symmetry d and the
equipotential q. For convenience write
= ^noC + Pa ) - Pm, <6'66)
where pri = PL/P m0 is the per-unit rotor leakage permeance, normalized

6-29
D esig n o f b ru sh less perm a n en t -m a g n et m o t o r s
to the base Then
u - a<l (6.67)

and when this is substituted back in equation (6.63) and rearranged,


= 2^ d - /r (6.68)
a d 1 + 2 P m R g lad

This is the total <axis flux produced by the armature reaction current
Its fundamental component is what generates the voltage in the
phasor diagram. The fundamental component is determined by Fourier
analysis of as described in [2]. Thus the amplitude of the
fundamental armature-reaction airgap flux-density symmetrical about the
d-axis is
jrP-P
?,ad = iIT Jf0
(6.69)

g
l ^ ud- V I -
The constants fc, and fclad arise in the Fourier analysis and are given hy
_4 an
v
and
r Sill OC7T
lad = a +
7r . ( 6 .7 1 )

If equations (6.67) and (6.68) are substituted into equation (6.69), there
results a simple equation for
*.ad - - ^ a d (6-72)
8i
where gd ' is the effective airgap presented to the fundamental
component of d-axis flux, given by

g" = __________ sL_________ .


, *i*d (6-7S)
lad i+

6-30
6. S inew ave M o t o r s
The use of the concept of the equivalent airgap makes it possihle to use
equation (6.54) for Xd, simply by substituting g instead of g : thus
+^ ^ i^ph )2 + ^ (6-74)
PSa
jfoie the addition of the leakage reactance Xo to the airgap component
V d to get the total reactance. For a two-phase machine the airgap
component Xad has 2/3 the value calculated by equation (6.74). The slot-
leakage reactance is calculated using methods described in Chapter 5.
for completeness, the fundamental armature reaction flux associated
with .Blad is given by
* = A ad ^stk (6.75)
* lad p
in [Wb]. The fundamental flux-ftn&jge associated with Jlad is =
iad and the peak induced voltage is G)ip]ad. The RMS induced
voltage is wt|lad/'/2 = ( 2 7 1 / ^ 2 ) ^ ^ /3> lad = Xdd/d V/phase.
The analytical derivation of equation (6.74) for Xd is essentially similar
io die procedure followed by Honsinger [1] and Miyashita [2] although
both these authors worked with pole-arcs of 180 electrical degrees (a =
1 ). The technique described here is similar to that of Richter [3-5].

6.4,2 Calculation of
The component of the stator ampere-conductor distribution associated
with / is symmetrical about the 9-axis and produces a flux that does not
pass through the magnet. Fig. 6.14 shows the ideal form of the airgap
ilux distribution. The airgap flux-distribution symmetrical about the <7-axis
i* given by
5 aq(0) = ^aq s <676)
g
and the total q-axis flux corresponding to is
SLTttlp

= 2*/ -^ la q ^Stk sin P dB


S (6.77)
^ II - cos 2 1 ] Flw
Pg 2

6-31
D e s ig n o f b r u sh l e s s p e r m a n en t -m a g n e t m o t o r s

This is the total 9-axis flux produced by the armature reaction current I .
Its fundamental component is what generates the voltage Xdq/ in the
phasor diagram. The fundamental component is determined by Fourier
analysis of 4> as described in [2]. Thus the fundamental armature-
reaction airgap'Vlux-density symmetrical about the 9-axis is
tt(2
8 (6.78)
= Alaq
k , Mac
F
g
where
*laq = IT (6-79)
Equation (6.78) can be used to determine Xq in the same way as
equation (6.54) was used for the symmetrical synchronous reactance X^
in the non-salient pole machine, and as equations (6.69) and (6.74) were
used for thus
xq = + = 6^ Lf + Xt (6.80)
pga
where
V = (6.81)
-*iaq
Note the addition of the leakage reactance Xa to the airgap component
to get the total reactance. For a two-phase machine the airgap
component X ,q has 2/3 the value calculated by equation (6.80). The slot-
leakage reactance is calculated using methods described in Chapter 5.
For completeness, the fundamental armature reaction flux associated
with filaq is given by
* laq = ffaq ^stk (6.82)

in [Wb]. The fundamental flux-iinAage associated with 5, is


*wl^ph^iaq anc* t^le Pea^ induced voltage is The RMS induced
voltage is wv|laq/%/2 = (2n/v/2)Awl7vrph / 1#q = X,q/q V/phase.

6-32
6. S inewave Moto rs

,4.3 Demagnetizing effect of d-axis flux due to 7d


Xhe d-axis armature-reaction flux 4*,^ produced by 7d is given by
equation (6.68). The proportion that flows through one magnet is given
by
^ = _ Pn _ X ^ = _ J ____ x f j d (6 8 3 )

^mo + 2 * + Pt\
where
Pa - (6-84)
and /*L is the leakage permeance per half-pole, i.e. the leakage
permeance in parallel with Pm0-
The corresponding flux-density produced in the m agnet is given by

B . 5 . (6.85)
where is the pole-face area of one magnet. This flux-density is
superimposed on the open-circuit flux-density in the magnet (equations
(4.10) and (4.13)). If 7d < 0, then flma < 0, i.e. the d-axis armature
reaction is demagnetizing, and the magnet operating point is driven
further down the demagnetization characteristic (Fig. 4.4).
6.4.4 Cross-magnetizing effect of q-axis flux due to 7q
The y-axis armature-reaction flux produced by I is given by equation
(6.77). This flux splits into two equal halves which flow across adjacent
pole-pieces. At the centre of each pole-piece (Fig. 6.14) the resulting
flux-density component is
* qad = (686>
This flux-density should be checked in design calculations to ensure that
the pole-piece has enough radial depth to avoid saturation.
6.4.5 Significance of rotor leakage
Rotor leakage has two main effects. First, it wastes magnet flux. On
open-circuit the airgap flux is less than the magnet flux. Accordingly the
6-33
D esig n o f brush less perm a nent -m a gnet m o t o r s

Fundamental Blaq
| Potential
jF la q ^ P 0

1 magnet
^mO
/> n
Zero potential
2p

Fig. 6.14 Geometry for the calculation of

leakage coefficient J^ kg *s defined in equation (4.5) as the ratio


in order to characterize the effectiveness of the magnetic circuit in ustfig
the available magnet flux. From the magnetic equivalent circuit, Fig. 4.7,

f = 1//?g = ------1----- (6.87)


1X0 PL + 1IRg 1 + PhRg
and
1 = 1 Pt R%, (6.88)
f,LKG
The second effect of rotor leakage is to divert d-axis armature-reaction
flux around the magnet, providing a bypass path. The per-unit rotor
leakage permeance pr] defined in equation (6.84) characterizes the
effectiveness of this bypass path. If PL is the rotor leakage permeance per
pole and ^MO is the magnet permeance per pole, then the relationship
between Jl&j ant^ PrI can ^ shown to be
6-34
6. S inew ave M o t o r s

A1 (6.89)

6.5 Phasor diagram


6. 5 / Non-salient-pole machines
If the magnets are on the rotor surface, and if the shaft cross-section is
circular, the sinewave motor is a 'non-salient pole 1synchronous machine:
that is, its daxis and ^-axis synchronous reactances are equal, = XQ,
and both are given by equation (6.55). In the steady state with balanced
sinusoidal phase currents, the operation can be represented by the
phasor diagram shown in Fig. 6.15. [Note that the squarewave motor
(Chapter 5) is not amenable to phasor analysis because the stator
ampere-conductor distribution is not sinusoidal and does not rotate.]
The d-axis is chosen as the real or reference axis and the open-circuit
voltage phasor E =jiL, where is given by equation (6.51).

'q

vd 'd d

Tig. 6.15 Phasor diagram of non-salient pole (surface-magnet) sinewave motor.


The phasor E l represents the voltage drop across the phase resistance R,
and is parallel to I. Similarly, the voltage drop across the synchronous
6-35
D esig n o f b ru sh less perm a nent -m a gnet m o t o r s
reactance is represented by \XJ, and leads the current phasor by 90".
The sum of the back-EMF and voltage-drop phasors must be equal to the
applied voltage at the terminals. Thus
V = E * (R + j Xt)I. (6.90)

6.5.2 Salient-pole machines


Fig. 6.16 shows a more general phasor diagram with the current and
voltages resolved into dr and f-axis components, and * X^, as in a
salient-pole synchronous machine. The inequality of and stems from
differences in the magnetic reluctance along the d- and ^-axes. In Fig.
6.16a the current leads the d-axis by an angle P > 90, and lags the ^-axis
by the angle y, sometimes known as the "torque angle", where
(6.91)

This equation applies also in Fig. 6.16fc, where y < 0 and the current lags
behind the ^-axis.
The current phasor is given by
(6.92)
where
1^ = /co s (3 = -/s in y; and (6.93)
Jq = /sin P = / cos Y-
Similarly the terminal voltage phasor V leads the ^-axis by the angle 6,
sometimes known as the load angle, and
(6.94)
where
V cos (5 + ) = - V sin S ; and
2 ( 6 .95)
V sin (5 + ) = V cos S .
2

6-36
6. S inew ave M o t o r s

JXJ
q'q

(a) (b)

Fig. 6.16 Phasor diagram of salient-pole sinewave motor in motoring mode, with
currents and voltages resolved into if- and ^ axis components, (a) rf-axis
current is 'demagnetizing'' (b) it-axis current is "magnetizing1'.

Also

Vi RI4 - XqIq and (6.96)


RIq * XAJd
and
V = E+ RJ+ jX dId * jX qfq. (6.97)
This equation reduces to equation (6.90) when Xd =
A positive d-axis component, /d > 0, produces an MMF distribution
around the airgap that tends to augment the d-axis flux produced by the
magnet. The stator current is said to be magnetizing. The flux
6-37
D esig n o f bru sh less perm a n en t -m a g n et m o t o r s

produced by the MMF associated with /d induces the voltage jXd/d in the
q-axis, which adds to ]E , as in Fig. 6.166. The magnet flux-density is
increased, so the operating point moves up the demagnetization
characteristic. It may even end up in the first quadrant with > Br
Conversely, a negative d-axis component, /d < 0, produces an MMF
distribution that opposes the magnet flux in the d-axis. The armature
current is said to be demagnetizing. The voltage jXd/d is in the
opposite direction to jq, as in Fig. 6.16a. The magnet flux-density is
decreased, so the operating point moves down the demagnetization
characteristic.
The power factor angle is given by
<f> = S - Y- (6.98)

Evidendy the power factor angle is decreased when 7d > 0. In other


words, a high power factor is associated with operation in the
demagnetizing mode, implying that a highly coercive magnet permits
high power factor and reduces the kVA requirement in the converter.
The electromagnetic torque is given by
T = 3p[i|rd/q - t|>q/d] (6.99)

in [Nm], where the flux-linkages ijtd and i|r are RMS per-phase values
given by
s'J'd = Eq * x d fd (6.100)

in [Vs]. The electromagnetic torque can therefore be written


r = ^ws [EqI + (Xd - X ) I dJq). (6.101)

The first term is called "magnet alignment torque" and the second term,
proportional to (Xd - Ar((), is called "reluctance torque". If 7d and are
substituted from equations (6.93) the torque equation becomes

6-38
6. S inew ave M o t o r s

T = &(J- E l cos y - ^2( X d - X H) sin 2Y1. (6102)

This shows that if Xd and X^ are unequal, the angle y required to


produce maximum torque per ampere varies with the current. Effectively,
the Aj- constant for the salient-pole type of machine is not constant but
depends on the current. This variation can be more complex than
indicated by equation (6.102), because the reactances Xd and Xq are
subject to saturation. In some motor configurations, notably the interior
permanent-magnet motor, this variation can be severe, especially in Xq.
If equation (6.102) is differentiated with respect to y, the angle for
maximum torque can be determined as

(6.103)
4 AX I

where AX = Xd - X . For example, if AX I = 0.2 , y = 10.7, and


therefore if AX I = 0.5^, y = 21.5.
[f = Xq, as in a surface-magnet motor, the torque equation becomes

(6.104)
This equation is the same as equation (6.53). A contour of constant
torque is therefore a horizontal line in the phasor diagram, with / =
constant.
6.5.3 Operation as a generator
fig. 6.17 shows the phasor diagram for generating, with both
demagnetizing (7d < 0) and magnetizing (7d > 0) orientations of the
stator current. It is usually more convenient to exchange the current
phasor for its negative, so that generating current appears positive when
leaving the machine, rather than when entering it: in other words, the
machine is treated as a source rather than as a sink. Then equation
(6,97) becomes

6-39
D e sig n o f bru sh less perm a n en t -m agnet m o t o r s

E = RZ+ j XAIA + jXq/q. (6-105)

Reversing the current in Fig. 6.12 causes the arrowheads on the HI, j-X ^
and jX^/ phasors to switch to the opposite ends of the respective
phasors. Ilie phasor diagram then appears as in Fig. 6.18, which shows
the usual case of lagging current. This is the classical phasor diagram for
an over-excited AC generator. The term "overexcited" generally means
that E > V. In other words, the excitation (in this case, the magnet flux)
must be increased above the value corresponding to rated terminal
voltage, in order to overcome the voltage drop in the synchronous
reactances and X^. The d-axis component of armature-reaction MMF
opposes the magnet flux.

Fig, fi.17 Phasor diagram of salient-pole sinewave motor in generating mode, with
currents and voltages resolved into A- and ^-axii components, (a) d-axis
current is "demagnetizing" (b) rf-axis current is "magnetizing".
6-40
6. S inewavk M o t o r s

fig. 6.18 Classical phasor diagram for generating wilh lagging load.

When PM machines are used as generators, the output is often rectified


in a diode bridge or a phase-controlled bridge. The DC output voltage
from the rectifier is given by
V* = ^ V ^ c o s a - l ^ c/dc ( 6 -106)
IT TT
where X, is the commutating reactance of the generator, a is the firing
angle (= 0 in a diode rectifier), and VLL is the line-line voltage at the
generator terminals. The commutating reactance is the reactance per
phase that inhibits or slows the commutation of the DC current between
SCRs or diodes in electrically adjacent branches of the rectifier. In a
plain PM machine the commutating reactance can be as high as the
synchronous reactance, which may be high enough to cause a significant
voli drop and loss of DC voltage. If the PM generator is fitted with a
retaining can around the rotor, the induced currents in the can
effectively reduce the commutating reactance to a much lower value,
comparable with the leakage reactance.

6-41
D esig n o f bru sh less perm a n en t -m a g n et m o t o r s

6.6 Circle diagram and speed/torquc characteristic


6.6.1 Non-salient-pole (surface-magnet) motors with
At a given speed is fixed by the magnet flux and the torque is
proportional to the ^-axis current Iq. Since E is itself proportional to
speed, this relationship is valid at zero speed also: see equation (6.52).
The linear relationship between torque and current simplifies the
controller design and makes the dynamic performance more regular and
predictable. It is characterized by the torque per ampere or torque
constant k^, which is discussed more fully in Chapter 7.
The amount of current that can be supplied is limited by the heat
dissipation capability of the motor, and by the current rating of the
controller. It is also limited by the maximum available controller voltage,
which must overcome both the back-EMF and the voltage drops across
the synchronous reactance and the resistance. In the complex plane of
the phasor diagram the maximum continuous controller current Ic limits
the current phasor to a circular locus described by the equation
yd1 2 + /q2 (6.107)
This is shown in Fig. 6.19.
The maximum available controller voltage Vc limits the motor terminal
voltage to a separate circular locus described by the equation
vd + K = vc- (6108>
If Vj and are substituted from equation (6.96), neglecting resistance,
then
X i q + {Eq X.I d)2 = Kc2 (6.109)
or
K ( 6 . 110)

This represents the circle labelled Voltage-limited current locus" in Fig.


6.19, with centre at the point (-o/^s 0> anc* radius (^c/ ^ - ^ ie
offset -q0/ i s independent of speed since both and ^ are
proportional to frequency, so the subscript 0 is added to denote a

6-42
6. S inew ave M o t o r s

V o lta g e -lim ite d


locus
q M ax c u rre n t
locus

In c re a sin g sp ee

Fig 6.19 Circle diagram for non-salient-pole (surface-magnct) PM synchronous


motor.
particular frequency such as rated frequency or corner-point frequency.
Operation within the controller limits must remain within the
intersection of the current-limit locus and the voltage-limit locus. If the
controller is a PWM sinewave amplifier, there is usually enough voltage
available at low speed to cause the voltage-limit locus to enclose the
current-limit locus so that maximum current can be obtained at any
orientation. The most advantageous orientation for the current phasor
is clearly the 9-axis, since this maximizes the torque per ampere, as
mentioned earlier. In general at low speeds the PWM duty cycle is low
and the phase voltage is chopped down to a value less than Vc.
Operation is along OQ, with torque proportional to current, and the
curront phasor /is as shown in Fig. 6.19.
As the speed and frequency increase, the current-limit locus remains
fixed, but there comcs a speed at which the radius of the voltage-limit
locus begins to decrease. This happens when the PWM duty-cycle reaches
Its maximum, and the motor terminal voltage equals the maximum
available sinewave voltage Vc from the converter. The PWM control is
said to be "saturated". Operation is possible along OQwith y = 90, but
643
D e sic n o f b r u sh less perm a n en t -m a g n et m o t o r s
it is only just possible for the current to reach its rated value Ic at Q. The
speed at which this happens is called the "corner-point" speed. It is the
maximum speed at which full torque can be developed.
If the speed increases further, the radius of the voltage-limit locus
decreases. The maximum current is at the intersection of the two circles,
such as P in Fig. 6.19. Although it is still possible to get the magnitude
of the current up to Ic, it is not possible to orient a current of this
magnitude along the y-axis and therefore the torque decreases. The
decreasing radius of the voltage-limit circle drags the maximum current
phasor further and further ahead of the y-axis, and the torque-producing
y-axis current decreases while the demagnetizing d-axis current increases.
This continues until the intersection P reaches point D, at which speed
the maximum current 7Ccan still be forced into the motor, but only just,
and it is entirely in the d-axis so that no torque is developed. The power
factor at this point is zero and the current is wholly demagnetizingan
onerous operating point for the magnet, especially at high temperature.
The loci OQ, and OD together form the limiting locus for the current
phasor throughout the whole speed range, and give rise to the
torquc-speed characteristic of Fig. 6.20. Along OQ maximum torque can
be developed with maximum current 7Coriented along the yaxis in the
phasor diagram. In physical space in the motor, the axis of the rotating
ampere-conductor distribution is 90 ahead of the rotor d-axis. Q is the
corner-point, the maximum speed at which full torque can be developed.
Along QD the torque decreases until at point D it is zero, with maximum
current 7C still flowing, oriented in the d-axis in the negative
(demagnetizing) direction.

Fig. 6.20 Torque/speed characteristic of non-salient pole sinewave motor


6-44
6. S inew ave M o t o r s
The ratio between the speeds at points D and Q is
k = .^2 = h .. ( 6 .1 1 1 )
UQ fQ
Jf we continue to neglect resistance, at Q we have
4 = 0; /q = /c; Vq = Eq0; Vd = - X ^IC. (6.112)

The subscript 0 in and in denotes values at the comer-point


frequency. From the phasor diagram
/ = j / q = j / c and Vc2 = Eq 2 + (6-HB)
so that
V2 - E 2
(6.114)
7c =

At point D,
> k - - 4 (6115)
"Tst)
Equating the two expressions (6.81) and (6-82) for 7C at the two
respective speeds,
V
k = ---------- c (6.116)
Z fi ~ J Vc ~ V
Suppose we define ^ as the per-unit open-circuit voltage at the comer-
point, with the maximum RMS voltage of the controller as the base
voltage:

^ (6-117)
Vc
Then

k = (6 .118)
" ft ~
6-45
D e sig n o f bru sh less perm a n en t -m a g n et m o t o r s

For example, suppose the motor is required to operate at maximum


torque up to 3,000 rev/min, and be capable of just reaching 6,000
rev/min with no load (zero torque). Then k = 2, and to achieve this ^
must be no higher than 0.911.
At the corner point the d-axis current is zero and, from the phasor
diagram at this point, the power factor is given by
cos 0 O = (6.119)
Kc
Also,
sin 00 = = (6.120)
*s() is the per-unit synchronous reactance at the corner-point frequency,
that is, the synchronous reactance normalised to a base of Vc/I c. In the
example quoted above with k - 2, the corner-point power factor is 0.911
and the per-unit synchronous reactance is xiQ = 0.411. Typically .v0 is
smaller than this in surface-magnet motors, giving a higher corner-point
power factor but a value of k closer to 1. The lower the per-unit
synchronous reactance, the lower the speed range that can be attained
above the corner-point speed. The surface-magnet motor has limited
capability to operate above its corner-point speed, and cannot maintain
a constant-power characteristic over a wide speed range. The
fundamental physical reason for this is that the airgap flux is fixed
predominandy by the magnet and direct field-weakening is not possible
to any useful degree. In this discussion the resistance has been neglected,
but in practice it introduces further phase shifts which make the
speed-torque envelope even more restriced than that presented. An
analysis that includes resistance is given by Leonhard [6 ].
If the speed is increased beyond point D in Fig. 6.14, there is a risk of
overcurrent because the back-EMF continues to increase while the
terminal voltage remains constant. The current is then almost a pure
reactive current (with nearly zero power factor) flowing from the motor
back to the supply. There is a small y-axis current and a small torque
because of losses in the motor and the controller. The power flow is
reversed and this mode of operation is possible only if the motor
'over-runs the controller. This can happen if the motor is driven by an
external prime-mover, or if the combined inertia of the motor and load
maintain a high speed after the controller voltage has been reduced. The
6-46
6. S inew ave M o t o r s
reactive current is limited only by the synchronous reactance, and as the
speed increases it approaches the short-circuit current E^/Xt. This can be
many times the normal continuous rating of the motor windings or the
controller, and may be sufficient to partially demagnetize the magnets,
particularly if their temperature is high. The current is rectified by the
freewheel diodes in the controller and there is a risk not only of
overcurrent in the diodes but also of overvoltage on the DC side of the
controller, especially if a filter capacitor and AC line rectifier are used to
supply the DC. Fortunately this operating condition is unusual, but in any
system design the possibility should be assessed. An effective solution in
cases of sustained over-running is to use an overspeed relay to
short-circuit the phase windings into a three-phase resistor or a
short-circuit, to produce a braking torque without stressing the
controller.
6.6.2 Salient-poU motors with Xd /
When Aj / X , equation (6.109) has the more general form
x j q (Bq * XdId)2 = vc2 (6.121)

with resistance neglected, and this represents an elliptical voltage-limited


current locus instead of the circular one in Fig. 6.14.

References
1. Honsinger VB [ 1982] Thr fteUh and parameters of interior type ac permanent magnet
machines, IEEE Transactions on Power Apparatus and Systems, PAS-101, 4, 867-575
I. Miyaahiu K, Yamashita S, Tanabe S, Shimozu T and Scnto H [1979] Development
of a high-speed permanent magnet synchronous motor, IEEE Transactions, Power
Apparatus and Systems
S Lafuze DL and Richter E [1976] A high-power rare-earth cobalt permanent magnet
generator in a variable speed constantfrequency aircraft starter-generator system, NAECON
conference record, 971-977
< Bailey LJ and Richter E [1976] Development report on a high-speed permanent-magnet
generator of the 200kVA rating class utilizing rare-earth cobalt magnets. Proceedings of the
second international workshop on rare-earth cobalt magnets and their applications,
Dayton, Ohio

6-47
D esig n o f bru sh less perm a nent -m a gnet m o t o r s
5 Richter E [1978] Trade-off studies for permanent-magnet machines tiling rar*-earth cobalt
magnets. Paper 1-4 of the third international workshop on rare-earth cohalt magnet*
and their applications, University of California, San Diego
6 Leonhard, W [19851 Control of electrical drives, Springer-Verlag, Berlin

6-48
7. kT and kE
7.1 Introduction1
The torque constant fej- and the back-EMFconstant A introduced in Chapter
1 are widely used to match the motor to its controller, especially when
the motor and controller are obtained from different sources. Even more
importandy, fcj. is used in the control-system design of complete servo
mechanisms that are driven by electric motors, because it represents the
essential "gain" of the motor in converting current into torque. Yet there
is scope for confusion in the way these factors are used. In S.I. units it is
often tacidy assumed that hj. and kL are identical. In English units, they
are numerically different, giving rise to the impression that they are
independent of one another. Sometimes kE is expressed in Volts per 1000
rpm, which seems at first sight to have no relation to the torque per
ampere.
The widely used Electro-Craft handbook [1) provides definitions for DC,
squarewave brushless DC, and two-phase sinewave brushless AC motors,
but it docs not discuss the three-phase sinewave BDC motor, neither does
it explain how these constants can be calculated from dimensions. The
scope for uncertainty is increased by two other factors :
1. Squarewave motors can be fitted with resolvers and operated
from sinewave drives. Indeed, in some manufacturers ranges the
differences between squarewave motors (brushless DC) and
sinewave motors (AC servo) are so small that the distinction
between them is not clear.
2. When applied to the three-phase sinewave motor, the "natural"
definitions of kj and kE used with commutator and squarewave
motors lead to a ratio of v^3/2 between kj and A. Ref. [1]
overlooks this factor. It only discusses the two-phase sinewave
motor, for which the problem does not arise.

There is some overlap between this chapter and other chapters that define Jfcj. and
kjr (Chapters 1,5 and 8); and also with Chapter 11 on measurement. The function of this
chapter is a detailed treatment of the two constants together with their measurement, in
which the similarities and differences among two- and three-phaae, and squarewave and
sinewave motors are brought out.
7-1
D esig n o f bru sh less perm a n en t -m a gnet m o t o r s

The definitions in this chapter arc designed to be consistent with the


usual methods of measuring Ap and kE, and these measurements are
described. In order to bring out the importance of kj and kE to the
fundamental energy-conversion process, losses are neglected throughout
this chapter.

7.2 Squarewave and sinewave motors


The definitions given in Chapter 1 are repeated a little more precisely.
A squarewave brushless DC motor is one which generates a trapezoidal
back-EMF waveform. It is operated from a DC supply with rectangular
line current waveforms. These waveforms are essentially direct current
which is commutated to the appropriate connection of motor phases, with
the appropriate polarity and conduction angle, by the controller
transistors. The transistors are also used to regulate the current
Squarewave motors can have any number of phases but 3 is a common
number. In this case, if the controller is a 6-transistor bridge, there are
normally two phases and two transistors conducting at any instant.
A sinewave brushless motor is one which generates a sinusoidal back-EMF
waveform. It is operated with sinewave currents which are normally
provided by a PWM inverter operating from a DC supply. Sinewave
motors can have any number of phases but B is a common number. In
this case, if the controller is a 6-transistor bridge, there are normally
three phases and three transistors conducting at any instant.
Squarewave motors commonly use Hall-effect position sensors to provide
the commutation signals. Sinewave motors usually use resolvers or
encoders (or their equivalent) to provide a continuous shaft-position
signal from which the necessary sine/cosine functions can be generated
to act as reference for the line currents.
Note that the terms trapezoidal" and "squarewave" are interchangeable
in this context

7-2
7. Kj. AND Kf
7,3 Definition and measurement of kj. and Ag
7J. 1 DC commutator motors
The DC commutator motor is the most basic of all electrical machines
from a control point of view. Viewed from the DC source, the ideal
brushless DC motor is electrically identical to the commutator motor.
The back-EMF constant ky and the torque constant form the basis of
Lhe control theory of both machines. To put this theory on a clear basis,
the equations of the DC commutator motor are summarized first.
The DC commutator motor is described by three simple equations as
follows:
T = Jfe, I Nm (7.1)
where Tis the electromagnetic (airgap) torque, Tis the DC current, and
Jtf is the torque constant in Nm/A. Next,
E = kE o m V (7.2)
where E is the back-EMF, com is the angular velocity in mechanical
radians/sec, and is the back-EMF constant in Vs/rad. Finally Ohms
Law: in the steady state
Ks = B * R^I Kb (7.3)
where is the supply voltage, RA is the armature resistance, and is the
combined voltage drop across both brushes. Many of the operational
characteristics of the DC commutator motor can be derived from these
equations, including the speed/torque characteristic and the variation of
current and voltage with speed and torque (see Chapters 1 and 5). For
the ideal DC motor,
kj = k 2 (7.4)
provided that hj. and are in consistent units (such as Nm/A and
Vs/rad). The basic energy conversion process is described by the
equation

7-3
D e sig n o f b ru sh less perm a n en t -m a g n et m o t o r s

T u m = - l E I = E l = Pelec Watts (7.5)


E

where -Pejcc is the electromechanical power conversion. It would be


convenient if all motors obeyed this equation, indeed if they obeyed aU
the foregoing equations.
Measurement. The obvious way to measure fcj. is to drive the motor as
a generator at a given speed and measure the open-circuit EMF. With no
current flowing, the brush voltage drop is zero and &E can be measured
direcdy as the ratio between the open-circuit EMF and the angular
velocity (in mechanical rad/sec). It is common practice to measure in
units of "Volts per lOOOrpm", but of course equation (7.4) is true only if
the units are V-s/rad.
An approximate way to measure is to run the motor from a supply
voltage Vt and measure the no-load speed. This method is approximate,
because the no-load current is not zero and therefore the terminal
voltage is not exacdy equal to the back-EMF E However, if the armature
resistance is measured in a separate test, and if the brush voltage drop
is known, then R J and can ^ subtracted from to give a more
accurate value for E~ This method is more convenient than the first
method because the motor under test does not have to be coupled to
another drive motor.
The torque constant fep can be assumed to be equal to k^, or it can be
measured in a separate test The motor is mounted on a torque table or
dynamometer, and load torque is applied from zero to maximum torque,
Fig. 7.1. If the torque is plotted against the current, the slope gives the
value of kj.. The torque should be corrected for friction and windage. At
higher currents Aj. decreases because of armature reaction, Fig. 7.2. (See
section 5.7.2).
7.3.2. Three-phase squarewave brushless DC motors
The definitions of fcj. and follow those of the commutator motor m
closely as possible, but the methods of measurement are slightly different
because the brushless motor can only operate as a motor when
connected to its controller. When driven as a generator it generates an
EMF waveform of alternating polarity, which is ideally trapezoidal.
7-4
7. ICj. AND Kj.

fig . 7.1 Measurement of fcj. of commutator or brushless motors

Torque T
s
/ ^ E f f e c t of
( armature
Jq- reaction
0 /
Friction C, Current I
torque

Fig- 7.2 Errors in measurement of Jfcj. can arise from armature reaction and
friction

Ajr can be measured bydriving the motor as agenerator and rectifying


theline-line voltagewith a rectifier having alarge smoothing capacitor,
Fig. 7.3a. The DC voltage across the capacitor is equal to the peak line-
line voltage <?LL. Then is defined as
kn = (7.6)

with units Vs/rad. At full speed, with no chopping, if the line-line back-
EMF waveform is trapezoidal with a flat top wider than 60, and if the
mean voltage drops in the transistors and the winding resistance are
small compared with the supply voltage, then
- Ks (7.7)

7-5
D esig n o f b ru sh less perm a nent -m a g n et m o t o r s

O n e phase

ph

Fig. 7.3 Measurement of of brushless motors: (a) 3-phase (b) 2-phase


where V, is the DC supply voltage. This shows that kF is equivalent to the
value defined for the commutator motor. The torque constant kj. can be
measured in the same way as for the commutator motor, as shown in Fig.
7.1, with the brushless motor operating normally on all three phases
from its controller. The current I can be taken to be the DC supply
current into the controller, or the peak line current. In ideal "brushless
DC" operation, these two currents are identical because the line currents
are rectangular waveforms 120 electrical degrees in width, with peak
value /, and two lines are conducting at every instant (In practice the two
currents are not identical because of commutation effects and because
the line current waveform is not perfectly rectangular.) Then
jfcr = I (78)
with units Nm/A. With these definitions, in the ideal case with no
saturation, no resistance, and no voltage-drops in the controller,
- (7.9)
and
Tut. 4 .i/ V
p
eiec
(7 .10)

7-6
7. K j-ANDK j .
where Pclcc has the same meaning as before, i.e. the electromechanical
power conversion. The ideal squarewave brushless DC motor thus has the
same basic energy-conversion equation as the DC commutator motor.
7.3.3 Two-phase squarewave brushless DC Motors
for the less common two-phase squarewave BDC motor the definitions
are similar and equations (7.6)-(7.10) can be used, except that the peak
line-line back-EMF e ^ is replaced by the peak phase back-EMF ph. In the
measurement of kE by the generator/rectifier test, only one phase of the
motor is used at a time, as in Fig. 7.36.
7.3.4 Two-phase sinewave brushless AC motors
The definitions of and fcp follow those given in Ref [1], Thus
(7.11)

with units V-s/rad, where ph is the peak phase voltage. This is measured
a
using capacitively-smoothed rectifier, as in Fig. 73b. If the back-EMF
waveform is sinusoidal, I h can be measured with an oscilloscope
connected direcdy across tne phase terminals; alternatively, an RMS AC
voltmeter can be used, and the reading multiplied by / 2.

The torque constant is defined as

(7.12)
with units Nm/A, where i is the peak phase current during normal
operation with two phases, i should be measured with a current-sensor
that is connected lo measure the phase current direcdy and display it on
an oscilloscope.
With these definitions, for the ideal two-phase brushless sinewave motor
J cj - kE (7.13)

7-7
D esig n o f bru sh less perm a n en t -m a g n et m o t o r s

and
7^ -*/ 2 ^ ^ - Pclec (7.14)
E
where Pclcc is the electromechanical power conversion, as before. In the
ideal case also
^elec = (7-15)
where I is the mean currcnt from the DC supply.
The equations of the two-phase sinewave brushless motor, and its
definitions of and ftp, are virtually the same as those of the three-phase
squarewave motor and the commutator motor, even though the
waveforms and principle of operation are quite different: the nvo-phase
sinewave motor is a classical rotating-field synchronous machine, whereas
the other two motors are DC motors, one being mechanically
commutated, the other being electronically commutated.
7.3.5 Three-phase sinewave brushless AC motors
Ref. [1] does not discuss the definitions of i E and ftp for the three-phase
sinewave motor. In fact there are several different ways in which they can
be defined. The one chosen here is selected so that
(a) the method of measuring is the same as for the other
motors; and
(b) the equivalent of equation (7.5) is satisfied in the natural
form for the three-phase motor (see equation (7.19)).
Thus *E is defined by
4 = (7.16)
with units V-s/rad, and it is measured by the circuit shown in Fig. 7.3a,
exacdy as for the squarewave motor, ftj. is defined in exacdy the same way
as for the two-phase sinewave motor, that is,

JcT = I (7.17)
T /

7-8
7. Kj. AND Kg
^rjih units Nm/A, where t is the peak line current when the motor is
operating normally, and is measured in the same way as described for the
fvvro-phase motor. However, with these definitions
(7.18)
and
^ Su.t An* v/24* (7.19)

where has the same meaning as before. The definitions of and


tj. are the same as for the two-phase motor, and easy to measure by the
same methods, but the three-phase sinewave operation makes kj * kE.
Since
^elec = K i (7.20)
in the ideal case,
Cll = vs and / = /. (7.21)
This shows that kj could be defined and measured as
(7.22)
when the motor is operating normally from its controller, and I is the
mean current taken from the DC supply. This alternative definition of kr
also applies to the two-phase motor, and may be more convenient for
measurement than the one in terms of peak phase current.
The ratio 73/2 between kj- and AE for the three-phase sinewave BDC
motor is an aspect of the "power invariance" principle. In effect, the
attempt to define and for sinewave motors is equivalent to Parks
reference-frame transformation in which the synchronous machine is
"transformed" into a DC machine. Parks transformation is given in
several alternative forms by different authors. A power-invariant
transformation is one which satisfies the equation
+ Vq = Va + + Vc (7.23)

7-9
d e s ig n o f bru sh less perm a n en t -m a gnet m o t o r s

where the symbols a,b,c refer to the phase voltages and currents, and d
and q refer to the transformed voltage and currents in d,q axes. This
equation expresses the principle that the instantaneous power measured
by the obvious direct method of summing vi products should produce
the same result in either the abc or the dq reference frame. But this i&
only possible if the reference-frame transformations contain appropriate
factors such as 3/2. The particular value (/3/2) of die ratio between kT
and arises because we have chosen to measure the peak line-line EMF
and the peak line current, rather than the RMS phase values.
7.3.6 Summary
The following table summarizes the definitions of kj and developed
in this section.

DC Squarewave Sinewave & Sinewave


3-phase squarewave 3-phase
2-phase

kf, Nm/A T /I T /\ 77ph T /\,

Ajj, Vs/rad E /u W m 4 l/m


in

At/* e 1 1 1 /3 /2

T able 7.1 S ummary o f Kj. a n d Kg f o r 2-phase and 3 - phase m o to rs

7.4 Calculation of ^ and ^


Although the methods for calculating the EMF of squarewave and
sinewave motors have been detailed in Chapters 5,6 and 8, it is often
desirable to calculate the constants &j. and k^ direcdy from the
dimensions and the magnet properties, and the necessary formulas art
7-10
7 . Kq, AND Kg
provided in this section, even though the material is essentially repeated
in different forms elsewhere in the book.
7.4.1 Squarewave three-phase brushless DC motors
The basic configuration is shown in Fig. 7.4. With the rotor rotating, the
F.MF ec generated in one full-pitch coil of Nc turns is
ec = NcB u mDL (7.24)
with units of volts, where B is the flux-density produced by the magnet
at the stator bore in the neighbourhood of the conductors of the coil, D
ii the stator bore diameter, is the angular velocity in mechanical
rad/sec, and L is the effective stack length. The EMF eph generated in a
whole phase winding is the sum of the EMFs ec for all th e coils in series
in that phase. If the number of turns in series per phase is N ^ , then
<Th = V ^ c - (?-25>

^8 - *7-4 EMF generation in a single full-pitch coil


7-11
D esig n o f bru sh less perm a n en t -m a g n et m o t o r s
The winding factor kw accounts for the fact that the flux-density is not
the same for all conductors. If we are talking about instantaneous EMF,
then in general varies with rotor position. However, if we are talking
about the peak, mean, or r.m.s. values of an EMF waveform, ^ is a
constant.
Now if a = PM/ 7t is the pole-arc/pole-pitch ratio (or per-unit pole arc) oF
the magnet, the airgap flux per pole <& due to the magnet is
nD L
2p
with units of Wb, where p is the number of pole-/w*r.j.
In the ideal wye-connected squarewave motor the waveform is
rectangular and two phases are conducting at any time. Therefore =
(where A means the peak value of the waveform), and if kE is defined
as the ratio <LL/(i)ra we get
Jc = (7.27)
n a
(For delta connection, the 4 is replaced by 2). An alternative expression
is sometimes used, in which Z, the total number of conductors in the
motor, is used instead of N^h. With Z = 3 x 2 x JVph, we get
t = (7.28)
E 3 an
(For delta connection, the 2/3 factor is replaced by 1/3). Usually the
conductors are sufficiently concentrated such that = 1 , but if a is too
small or the coils are too short-pitched or skewed, will be less than I.
If there are a parallel paths in the winding, the expression for is
reduced by the factor a.
The torque constant kj- is the ratio T/I, where T is the electromagnetic
torque produced by two phases carrying current /. Assume that the
windings are sufficiently concentrated so that each phase produces
uniform torque for an interval of 120 electrical degrees. Then
T = * B IL (7.29)
3 2

7-12
7. Kp AND Kj.
and if we substitute for from equation (7.26) we get
k, - \ k . (7.30)
3 air
which is exactly the same as equation (7.28), showing that ftp = for this
ideal motor. If there are a parallel paths, the expression for ftp is reduced
by the factor a. The corresponding waveforms of line-line EMF and
current are shown in Fig. 7.5. The equality of AE and ftp holds regardless
of the connection (wye or delta) or the number of parallel paths.

Fundam ental Trapezoidal EM F

60 JL^L

T _L

Fig. 7 .5 Waveforms for 3-phase squarewave motor


7.4.2 Two-phase sinewave brushless AC motors
The generated EMF of one phase has the peak value
4* = * r ' K l * b * i = wm P K \ (7-31>
where fewl is the fundamental harmonic winding factor, and $>j is the
fundamental flux/pole at the stator bore due to the magnet. includes
7-13
D esig n o f b ru sh less perm a nent -m a g n et m o t o r s
the distribution, pitch, and skew factors of the actual winding: i.e. ^ =
*dl* pl* sl- W ith kr = ^ p h /^ n H
K *H 1V 1 (7 -32)
To derive the torque constant we need to determine the torque
produced by the interaction between the rotating ampere-conductor
distribution and the magnet flux. In the ideal sinewave motor, only the
fundamental space-harmonic components need be considered. The
rotating magnet flux at the stator bore is distributed according to
B( 8) = Bx cos (p 8 - Q t) (7.33)
where CO = 2nf, and the fundamental flux/pole is

= B \DL (7.34)
P
The actual winding produces an MMF distribution whose fundamental
component is the same as that of a concentrated winding having &wl jV)h
turns. This is equivalent to a sine-distribution of C\cos />0
conductors/radian where
c, -7T2-A ,,A r1*. <)
The other phase winding has an equivalent sine-distribution of
conductors sin pft. With currents * cos oz and * sin to* there is a
rotating ampere-conductor distribution
Cl /cos pB cos ai t + C J sin p 6 sin o> t
1 1 (7.36)
= C,/cos {p 8 - (i>/)
and the resulting torque is

T= 2p f j lp CliB l ~ L cos2 ( p9- ut ) d 8 . (7.37)


If we complete the integral and substitute for /{, and from equations
(7.34) and (7.35) respectively, we find that

7-14
7. KpANDKj.

T = ( /^ A ^ ,) ! = V- <738>
By comparison with equation (7.32), we see that fep and kF are identical.
If there are a parallel paths in the winding, the term in brackets in
equation (7.38) is divided by a.
With two phases, if the MMF waveform is oriented so that its axis is
orthogonal to that of the magnet flux waveform (i.e. if the current is
oriented in the q-axu), the internal power-factor angle between the
generated EMF phasor E and the current phasor I is zero and
r o m = k, /x j k = h S i- 2 (7-39)
as in equation (7.14).
7.4.3 Three-phase sinewave brushless A C motors
In Che three-phase motor, assuming balanced sinusoidal operation and
wye connection,

and therefore from equations (7.16) and (7.31) we get


kE - 73 <7-41>
with units V-s/rad. The fundamental rotating ampere-conductor
distribution isderived in a similar way to equation (7.36), but because
there are three phases whose axes are displaced by 120 electrical degrees,
with three-phase balanced currents, the result is [2]
C,/cos (p 6 - <iii) (?-42)
and when this is integrated with the rotating magnet flux distribution the
torque is found to be
t - * = V - <7-43)
Hence
f <7 44 )

7-15
D e sic n o f b r u sh less perm a n en t -m a gnet m o t o r s

as in equation (7.18). If there arc a parallel paths in the winding, the


expressions for kj and are divided by a. If the winding is in delta, the
expression is divided by two.

7.5 Example calculation


Suppose we have a brushless motor with a set of line-line EMF waveforms
as shown in Fig. 7.5. The peak line-line EMF is = 60V at lOOOrpm,
giving
JtE = ------ ------ = 0.573 V-s/rad.
1000 *
60
For such a squarewave motor kj- = k^, so if the rectangular current
waveforms have a peak value of 4 A, the torque is 0.573 x 4 = 2.29Nm.
The RMS phase current for this waveform is /(2 /3 ) x 4 = 3.27A.
This motor could be operated as a sinewave motor with sinewave
currents. The fundamental component of the generated line-line EMF
can be determined by Fourier analysis of the trapezoids in Fig. 7.5, and
the result is = 1 053 x 60 = 63.2V, giving = 0.603Vs/rad. It is
remarkable how close the fundamental is to die peak value for a
waveform of this shape.
From equation (7.18), fcj. = 0.522Nm/A. The current can be chosen in
two ways. First, if the peak current is the same as in the squarewave case,
i.e. 4A, then from equation 17 the torque is 2.09Nm, which is 91% of the
torque produced by the squarewave drive at the same peak current
However, the RMS current is only 4/^/2 = 2.83A.
Alternatively, if the same RMS current is used, the peak sinewave current
is i/2 x 3.27 = 4.62A and the torque is 4.62 x 0.522 = 2.41 Nm, which ii
105% of the torque produced by the squarewave drive.
References
1. Electro-Craft Handbook, Fifth Edition, August 1980, ISBN 0-960-1914-0-2
2. TJE Miller [1989] BnuMftt prrmnjicni-magnet and nlueUma motor draff, Oxford
University Press, ISBN 0-19-859369-4

7-16
8. THE BACK-EMF WAVEFORM
S,L INTRODUCTION
"This chapter is concerned with methods for the rapid calculation of the
back-EMF waveform, including the methods used in the design program
pC-BDC (see Chapter 13 and Ref. [1]). A good estimate of the back-EMF
vaveform is required for two main reasons. In the first place, it is an
important indicator of the ability of the motor to produce smooth
torque. Squarewave motors need a back-EMF waveform that is essentially
flat throughout the commutation zone, while sinewave motors require
essentially sinusoidal back-EMF waveforms. In theory it is possible to
profile the current waveform by chopping so that smooth torque is
produced with any back-EMF waveform, but this requires sophisticated
electronics and is only occasionally proposed in practice [2]. Most of the
time, designers aim to get the back-EMF waveform as close as possible to
the trapezoidal or sinusoidal norms, within reasonable limits imposed by
manufacturing constraints.
The second reason for requiring a good estimate of the back-EMF
waveform is for accurate simulation of the motor operating with its
controller, and for the determination of the current waveform and the
correct control strategy.
We have seen in Chapter 7 that the usual formula for the peak line-line
back-EMF of a 3-phase, wye-connected squarewave brushless DC motor
operating with "two phases on" is
^LL = ^EWra
where is the back-EMF constant given by equation (7.28). These
formulas give no information about the EMF waveform, which is generally
assumed to be trapezoidal or sinusoidal. However, when the designer is
making basic choices about the slot number, pole arc, slot opening, coil
pitch, and other dimensions, he or she may have no a priori rules which
guarantee that the back-EMF waveform will have an acceptable shape. It
is therefore desirable to have a quick means of calculating it, with all the
key dimensions and parameters taken into account.
The calculation of the whole back-EMF waveform is inevitably more
complicated than the calculation of ftE or hj, because the values of these
8-1
D esig n o f bru sh less perm a nent -m a g n et m o t o r s

constants depend only on the total flux, and when they are calculated ft
is usually assumed that the EMF waveform is of the correct shape.
However, we have already seen in Chapter 3 that when the basic choices
of slot number, pole number, pole arc, coil span, etc., are being made,
there are no a priori rules which guarantee the correct EMF waveshape.
The process of selecting the correct values of these parameters needs to
include a visual inspection and perhaps also a harmonic analysis of the
EMF waveshape for each trial design. This requires a computer program
because the number of variations is far too large to be comfortably
handled by manual calculation and graph-plotting. Moreover, the
influence of fringing is very' difficult to calculate by hand. Even though
methods exist, such as conformal transformation, a computer is still
required to perform the numerical computation of the results, and to
work through the variations of key parameters such as magnet length,
slot opening, skew, and other factors.
Methods for calculating the back-EMF waveform can be classified into
analytical and numerical methods. The main analytical methods are (1)
the lumped-parameter reluctance-network (magnetic equivalent circuit);
(2) conformal transformation; and (3) analytical solution of Laplace's
equation. The main numerical methods are (1) the finite-element
method and (2) the boundary-element method, both of which are
available in well-developed computer software packages.
What is easily overlooked is the amount of numerical calculation involved
in design, which is an iterative process involving typically dozens or
hundreds of parameters, all of which have to be varied over quite wirir
ranges and finally fine-tuned to get an "optimal" design or an acceptable
compromise. Not all of these are electrical or magnetic: they include
thermal and mechanical parameters and calculations. Much of the
current research and development in motors is overly concerned with the
use of ever more sophisticated techniques to analyse one particular design:
this is especially characteristic of publications relating to the iinite-
element method. But the finite-element method is orders of magnitude too
slaw and expensive for use in normal design work. For the present, and
probably for some time to come, design calculations will be found to be
more productive and more efficient when they are based on analytical
methods which are much faster, even though they may be less accurate.
The proper use for finite-element analysis is the checking and refinem ent
of designs, or the analysis of problems that are simply too difficult for
analytical techniques.
8-2
8. T h e Back -EM F W aveform
f\ n elegant and effective example of the application of computerized
analytical methods to the EMF waveform calculation is the work of Boules
[3,4]. Boules method is based on the representation of the magnets by
equivalent current-sheets, using a formulation that dates back to Hague
[5|. Unfortunately the technique does not account for the effects of slot-
opening geometry or skew, and it is limited in the rotor configurations
which can be dealt with. Yet several current designs of commercial
importance use a small number of slots per pole (including fractional
values such as 1.5), and many brushless motors are skewed, either in the
stator or in the magnets. On the other hand, Boules method is ideal for
analysing slotless motors (Chapter 5).
The finite-element method is possibly the most powerful tool for
analysing the effects of slotting, as has been demonstrated in the
extensive papers of Demcrdash and his colleagues [6 ]; but it is too
cumbersome for the present purposes, where virtually instant results are
called for, and it is arguable whether it is yet a practical tool for
modelling the effects of skew.
Most authors have treated the back-EMF calculation and the core-loss
calculation as completely separate exercises, without mentioning the fact
[hat in a comprehensive design theory the two calculations must be
absolutely consistent with each other. Ref. [7] combines both analytical
and finite-element models of the EMF waveform and iron losses in
brushless machines, and goes some way towards answering this criticism.
In this and the next chapter it is shown how the core-loss calculation can
be made consistent with the EMF waveform calculation, when both are
based on the tooth Jlux waveform.
The theory in this and the next chapter is based on relatively simple
magnetic field computations modified by smoothing or fringing functions
which, in the PC-BDC computer program, may be adjusted or turned on
or off at will, giving physical insight into the effects of the design and
layout of the slots, magnets, and windings on the back-EMF waveform,
core losses, and performance. The fringing functions are simple
analytical functions, based on only a few points on the waveform.
This chapter describes the back-EMF calculation by three alternative
procedures, indicating their appropriateness for different situations. The
simplest is the "BLV" method, which is the most widely used. A discussion
of its limitations leads to two methods based on the tooth-flux waveform,
8-3
D esign o f bru sh less perm a nent -m a cn et m o t o r s

and this provides the necessary link with the core loss calculation in the
next chapter.

8.2 The BLV method


The "BLV" method treats the stator conductors as filaments on the
diameter D on the bore of a smooth, unslotted stator. The waveform of
EMF in any coil originates from Faraday's law:
* - - 4i 4t (8.2)
where i|r is the flux-linkage of the coil [V-s]. Allowing for variations in i|j
with time and rotor position , the EMF can be written
e = 31 + i t M = at
dt dt d/
The "transformer EMF" term 3i|t/dt disappears, provided that the
reluctance of the magnetic circuit remains fixed, so that the magnet flux
does not vary with time. Suppose the coil-sides are at angular positions
and 02, so that the coil pitch is 0 2 - 0r Then i|r is proportional to the
integral of the magnet flux-density distribution 5(0) around the airgap
between 0^ and 02:
t(r = Tcr t f 2 B(6)d9 <8-4)
where Tc is the number of turns, r = D/2 is the bore radius, L is the stack
length, and 0 is the azimuthal angle around the stator bore. Equation
(8.4) should be substituted into equation (8.3) to give
e = 0)mTztL * f 1 B O )* . (8-5)
The differentiation under the integral sign can be simplified only if
and 0 are equal or differ by a constant, meaning that the flux-density
distribution 5(0) is fixed to the rotor and does not vary in shape as rln
rotor rotates. In that case the result is
e = {B1 - B^LV

8-4
8. T h e B ack -EM F W aveform
vfhere V= rtom is the linear velocity at the stator bore and and are
die flux-densities at 0j and 02 respecdvely. In a full-pitch coil, further
simplification results from the fact that B^ = -B^, giving e = 2 BLV.
In the "BLV method 5(0) is the static magnet flux distribution around
the airgap, unmodulated by slotting. According to the "BLV" method the
EMF waveform in a full-pitch coil has the same shape as the unmodulat
ed airgap flux-density distribution. In a surface-magnet motor this has the
form shown in Fig. 8.1.

Fig. 8.1 Airgap flux-density distribution with smooth stator

8.3 Airgap flux-density distribution


The underlying structure of the flux-density distribution j5(0) in Fig. 8.1
is a rectangular pulse of width pM, the magnet pole arc in electrical
radians (PM = an). The origin is the ^-axis or interpolar axis, and the
hreakpoint angles 0a and 0b are then given by
. ^ ; f t - *. * ft,- <8-7>
The fringing outside PM is represented by the single smoothing function
l e-*a (8.8)

8-5
D esig n o f bru sh less perm a n en t -m a g n et m o t o r s

where x= - on the leading (left-hand) side and x = - on the


trailing side. Within pM the smoothing function is
1 (1 - e~*) ( 8 .9 )

where x = f; - a on the leading side and * = b - \ on the trailing side.


The fringing exponent a can be estimated by

( 8.10)
g /r>
where Lm is the magnet length and |J.rcc is its relative recoil permeability.
This formula, which was proposed by Dr. R Rabinovici [8], has its roots
in Hagues analysis of the unslotted motor with L^ = 0 [5], and is found
to give good results in practice (see below). The total width of the fringe
zone on each side of the magnet is taken to be fa = 7a, a range which
encompasses 97% of the exponential transition. The significance of the
product term in equation (8.10) is that fringing vanishes if either g or
(Lm + g) is zero; both of these conditions agree with physical intuition.
The N and S magnet poles are treated independently according to
equations ( 8.8) and (8.9), and the resulting EMF waveforms are added.
If the magnet arc is very narrow, the fringing function from the left can
overlap with the one from the right. In the overlap region the resultant
is taken to be
y = i - y% - yr (8-n )
where y2 is the fringing function from the left and y3 is the one from the
right. The justification for this is explained in Fig. 8.2. A narrow magnet

Fig. 8.2 Overlapping fringing functions - narrow magnet pole arc


8-6
8. T h e B ack -EM F W aveform
N can be replaced by five magnets N, N2, NS, S2 and S3. The three
North magnets combine into one wide magnet, so that the total flux-
density in the centre contributed by the North poles is 1. The flux-
densities contributed by the two South magnets are respectively and
giving a total 1 - - y3 within the arc of the narrow tooth.

g,4 Skew
A skewed magnet can be considered as being made up of many small
magnets arranged as in Fig. 8.3. The flux 4>T passing through the whole
axial length of one tooth is nearly the same as if this array of magnets
was replaced by a stack of thin sheet-magnets, each having the full length
/.in the axial direction, but with graded arcs. The combined strength of
the sheet magnets is such as to produce the same airgap flux distribution
as the actual magnet.
If the fringing functions of equations (8.8) and (8.9) are applied
individually to each sheet magnet, while their thicknesses are allowed to
decrease to an infinitesimal value, the total effect can be obtained by
integrating the field contributions of the sheet-magnets over the range
-o/2 to o/2, where o is the skew angle as shown in Figs. 8.3 and 8.4. If
the contribution of one infinitesimally thin sheet-magnet to the total flux
distribution is dB = (Bpk/o)d0j, then in region 1 of Fig. 8.4 the total is
given by the integral over all the sheet magnets, i.e.,

5,(0) = C dBQ) = r 3 * X 1 e(e e')ladBl


J -ofl J -o/l o 1
( 8 .12 )

Similarly the flux-distribution in region 2 of Fig. 8.4 is obtained as


Btf) = ^{1 - i -L [ _ a-9*mn ]} (8.13)

and in region 3 as
B^B) - q, k{ " 7 ~ (8.14)

8-7
D e sig n o f b r u s h le s s p e rm a n e n t-m a g n e t m o t o r s
L L

Fig. 8.3 Effect of skewing on airgap flux-density distribution

Fig. 8.4 Integration of the effect of skewing

These equations approximate the equivalent airgap flux-density waveform


including magnet-fringing and skew, in a fonn that is well adapted for
rapid computation.
8-8
8. T h e B ack -EM F W aveform
fhe equivalent sheet-magnets mask the effects of any variations in the
jnajjnetic field along the z-axis. In reality, axial fringing changes the
^agnet operating point slightly, but this effect is ignored. The teeth and
conductors integrate any z-variation to give a single total tooth-flux (or
conductor EMF) at any rotor position

8.5 S lo ttin g

Slotting causes difficulties with the "BLV method. It is not obvious how
the EMFs in conductors located in slots are related to the EMFs in
filaments on the bore of the unslotted stator. Moreover, slotting
modulates the airgap flux-distribution ij(0) at the stator bore, Fig. 8.5.
The modulating function is fixed to the stator while the fundamental flux
rotates with the rotor: the 5(0) waveshape changes as the rotor rotates,
and H and 0 in equation (8.5) are not simply related to each other.
Evaluation of e therefore requires the prior evaluation of the flux-linkage
integral as a function i|r(E) that can subsequendy be differentiated with
respect to

fiij- 8.5 Airgap flux-density distribution -effect of slotting


To evaluate the function i|f()> (e-g- by a series of finite-element
calculations, many values are required covering a complete pole-pitch of
rotor rotation. The resulting waveform ijr() is unique to the values of
0j and 02, and the calculation takes a long time. It therefore makes sense
to select these coil-side locations such that the entire phase-EMF can be
reconstructed by superposition of a set of phase-shifted i(r( 0 functions
from all the coils of the phase winding. In conventional machine analysis

8-9
D esig n o f bru sh less perm a nent -m a g n et m o t o r s
such a building process is based on the full-pitched coil, i.e., 10 j - 02|
7T electrical radians, but not all windings can be decomposed into full,
pitch coils. A better choice is the single-tooth coil, which has the
following advantages:
1. It is more general in the sense that any winding can be
"decomposed" into an electrically-equivalent series of single-tooth
coils. Some motors are in fact wound with single-tooth coils.
2. The tooth flux-density is approximately proportional to the flux
<f>T linking a single-tooth coil, and its waveform can be used to
compute the core losses in the teeth more rigorously than with
conventional core-loss formulas.
3. The flux waveform <J)y in any yoke section can be reconstructed
from the tooth-flux waveforms. This can be used to compute the
core losses in the yoke, and it is also useful as an indicator of
cogging torque.
4. Single-tooth search coils are convenient for measurement, and
they can often be fitted after the stator is wound. It is more
difficult to fit full-pitch search coils particularly if they have several
turns.
5. Analysis based on the single-tooth coil is particularly useful
when considering stator saturation conditions with the machine
loaded.
Notwithstanding its limitations, the "BLV" method is often satisfactory
particularly if the slot openings are narrow. It has the advantage of being
simple and fast. Although not every phase winding can be decomposed
into full-pitch coils, the total phase EMF can be reconstructed from the
totality of conductor EMFs.1 It can also deal with skewing if the airgap
flux-density waveform is modulated according to equations (8.12-8.14).

1Because the fringing flux enters the sides of the teeth the effective slot openings are
considerably narrower than the actual ones. This effect is probably more pronounced in
surface-magnet PM motors than conventional nanow-gap machines because of the larg
effective airgap through the magnet.
8-10
8. T h e B ack -EM F W aveform
8.6 Calculating back-EMF from Tooth Flux
8.6.1 Single-looth flux and EMF
The single-tooth EMF is the back-EMF in a coil wound around one
tooth, and is given directly by Faradays Law:
e. = (8.15)
^ dt m^
where <j>T is the tooth flux due to the magnet (with no current flowing
in the stator), and (i)m is the angular velocity in mechanical rad/s. In
terms of the tooth flux-density, is given by
e, - (8.16)
where B j is the flux-density averaged over the effective tooth area Ar and
f; is the rotor position, is the tooth width; and ks is the lamination
stacking factor. In this analysis, variations of flux-density across the tooth-
width are ignored, and the flux through the tooth Is assumed to vary only
as a result of the rotation of the magnet.
As the edge of the magnet sweeps across the tooth surface, the tooth flux
<j>T varies as shown in Figs. 8.6 and 8.7.
Fig. 8.7 shows the idealised variation of <J>T with no fringing or skew. It
is easier to understand if the magnet is considered fixed while the tooth
sweeps across it from position 1 through position 5. At position 1 the
tooth flux is zero. It reaches its maximum value aT radians later at
position 2, where the whole tooth overlaps the magnet. After that, from
position 2 through position 4, there is no variation in the tooth flux.
Finally, between position 4 and position 5 the tooth flux falls to zero.
Because of the fringing at the edges of the magnet, the "transition angle"
between zero and maximum values of <}>,. is wider than the actual tooth
arc. It is as though the tooth arc a T were augmented to a larger value
a-pp. The transition essentially takes place over the augmented tooth arc
ar r In Fig. 8.6 the augmented tooth arc would be nearly equal to the
slot pitch. This supports the assumption on which the "BLV" method
depends, namely that the slot openings are negligible; however, the slot
openings in Fig. 8.6 are relatively narrow. The maximum tooth flux is

S-ll
D esig n o f bru sh less perm a n en t -m a g n et m o t o r s

Fig. 8.6 Fringing around the edges of tooth and magnet

clearly gathered from the augmented rather than the actual tooth arc.
A possible formula for augmenting the tooth arc is
rr = ys - K wo (8-17)
where is the augmented tooth arc, yt is the slot pitch, and is the
slot opening (all in electrical radians). The Carter coefficient kQ is
obtained as a function of the ratio w0/g-from Carters graph [9]. For the
surface-magnet PM motor, we replace g by the effective airgap g +
V ^ re c
The tooth EMF is the derivative of the <t>T waveform in Fig. 8.7, and
the idealised trapezoidal <j)T waveform therefore produces rectangular
pulses of back-EMF. The next two sections describe two alternative means
for determining the <t>T and waveforms.

8-12
8. T h e B ack -EM F W aveform

Fig- 8.7 Accumulation of tooth flux as the rotor rotates. The back-EMF waveform is the
derivative of the tooth-flux waveform

8.6.2 Accumulation of tooth flux


The tooth-flux waveform can be obtained by numerical integration or
"accumulation". Fig. 8.8 shows the notion of a magnet pole sweeping
across a skewed tooth. The tooth flux waveform is generated by
accumulating flux at the overlap line, at a rate determined by the airgap
flux-density distribution of Fig. 8.1. In the PC-BDCcomputer program the
tooth-flux accumulation algorithm steps the rotor in increments of 0.5
electrical degrees. If the tooth is skewed, the width of the overlap line
varies as shown in the lower diagrams of Fig. 8.8, and these functions
modulate the rate of flux accumulation. These modulating functions have
the same form as Fig. 8.4, and it can be shown that this procedure is the
same as assuming a straight tooth and representing the magnet (with the

8-13
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

C tfC V A / A N i m F a T O O TH T2

L2

0<rr MAGNET

o Li v l-c - -
(a) Low skew (b) High skew

Fig. 8.8 Tooth flux accumulation with skew, (a) Low skew (b) High skew

same skew angle) by the flux-distribution functions in equations


(8.12-8.14).
An advantage of the tooth flux accumulation procedure is that in
software, it is easy to accommodate bifurcated teeth, Fig. 8.9.

Fig. 8-9 Bifurcated teeth

8.6,3 Direct construction of tootk-EMF waveform


Differentiating the tooth-flux waveform cf>T to get the waveform tends
to aggravate any imperfections in the approximating functions especially
8-14
8. T h e B ack -EM F W aveform
if these are piecewise. Recognizing that the same "fringing zone and
"breakpoint angles" underlie both waveforms, it is of interest to examine
whether can be approximated first, and subsequendy integrated to get
the <t>T waveform.
The underlying structure of the er waveform is a rectangular pulse of
width a .^ , Fig. 8.7. This funcdon can be smoothed outside by the
single function
(8.18)
where * = \ j - on the leading side and x = {; - on the trailing side,
and E is the rotor position, { being the rotor position labelled 1 in Fig.
8.7. Smoothing or fringing within is represented by the function
(8.19)
where x= I; - Ej on the leading side and x = E2 - on the trailing side.
The fringing exponent a is again estimated by equation (8.10),
underlining the fact that the fringing is predominandy related to the
magnet flux rather than to the slot geometry. If the tooth arc is narrow,
the principle employed in Fig. 8.2 is applied to merge the left and right
fringing functions.

The ffj. waveform as described hitherto is the result of one edge of one
magnet passing one edge of the tooth. The other edge of the magnet
passing the other edge of the tooth produces another EMF pulse whose
edge is shown dotted and inverted in Fig. 8.10. Where these waveforms
overlap, they should be added together, producing the resultant eTS (for
the south magnet pole) which passes through zero as shown. A similar
image exists, folded about n, and the same addition must be performed
with it.
The north magnet produces an EMF pulse ,?TN identical to that of the
souLh magnet, but displaced symmetrically about the angle n / 2 and
shown in Fig. 8-10 as The total tooth EMF is given by
^TS + ^TN' ( 8 .20)

The overlap between EMF pulses coincides with the region of minimum
tooth flux, implying that the superposition is valid even though the tooth
may be saturated at other rotor positions.

8-15
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a c n e t m o t o r s

RESULTING WAVEFORM ;
FROM S MAGNET

LEADING
EDGE

TRAIUNG
EDGE

Fig. 8.10 Construction of tooth-EMF waveform by superposition of contributions


from the N and S magnet poles

It is convenient to normalize the EMF waveform functions to the


maximum value which is given by
= CJm^ T. (8-21)
a TT

BTp is the peak flux-density in the tooth when it is aligned with a magnet
axis. The use of the waveform functions means, of course, that this
magnetic field calculation is performed at only one rotor position an
important factor in the overall economy of the method. The correct
value of the phase EMF, and in particular its peak value, depends on an
accurate calculation of the peak value of Bj. The quickest method for
this is a nonlinear reluctance network in which the teeth, yoke, magnet,
and airgap are modelled as series elements, with appropriate allowances
for magnet leakage flux, as described in Chapter 4. Alternatively, the
magnet can be represented by an equivalent surface-current distribution
provided that its relative recoil permeability is close to unity. Of course,

8-16
8. T h e B ack -E M F W a v efo r m
finite elements can also be used to calculate Bj accurately but with a
little more effort.

8.7 Construction of phase EMF from the e-p waveform.


The complete phase EMF can be obtained from a single tooth EMF by
decomposing the phase winding into a set of electrically equivalent
single-tooth coils. Consider a single full-pitch coil wound in slots 1 and
10 which is electrically equivalent to a set of nine single-tooth coils
wound in slots 1-2, 2-3, 3-4, 9-10. The EMF waveforms and flux
waveforms <j)T of these coils are identical, but each is phase-shifted by the
slot-pitch angle from the previous one. This is shown in Fig. 8.11.
Fig. 8.11 shows the rectangular pulses that would be obtained with no
fringing and a very wide slot-opening [8]. When the fringing is added,
the resultant is smooth. In the PC-BDC computer program, the
construction of the phase EMF is performed automatically from the coil
table, which contains com plete information about the spans, numbers of
turns, and connections of all coils in the winding.

SINGLE-TOOTH EMF WAVEFORMS (NO FRINGING)

S m agnet 2s Jl___i[
3S j l Z j v
"U Lf3N
4S_n__ n_
u U"4N
5S_n__ n_
U L f5 N
6 s_n___ n
U U" 6N
7S_n__ n.
U Lf 7N
8S j i ___ n_
u LT0 N
9S j i __ n_
U----- LT 9N
d im m
SMOOTH WAVEFORM
WITH FRINGING
Full-pitch coll EMF
Coll-pltch = 9 Blots

fig. 8.11 Conjunction of full-pitch coil EMF from tooth EMFb


8-17
D esig n o f b ru sh less perm a n en t -m a g n et m o t o r s

8.8 Development of yoke flux waveform from 4>T waveform


The yoke fluxes j = I - 71 can be represented together with the looth
fluxes (J)j as in Fig. 8.12. Evidendy
0j = j] - y\*i. j = 1,2,-jV ( 8 .22 )
where N is the number of teeth, and
y\ + y-i+ - + yN = - (8.24)

From equation (8.22),


yi = y\ " ^1
= > 2 - 0 2 = 7i - 01 02 (8.24)

7a = / i " 0i ' 02 - - - 0*
Adding equations (8.24),
Yi + (Si - 0i) + Oi - 0i - 0j) + - (8.25)
+ (Ti " 0i - 02 " - - <f>t) = 0
fro m w hich

8-18
8. T h e B ack -EM F W aveform

y \ = 4> ( * - i) * j- (8 -26)
" j=l
This means that the yoke flux in any section can be constructed as a
weighted sum of all the tooth fluxes. If the number of slots per pole-pair
is even, the expression simplifies to
*2
A - \ Ej=i *i- (8-27>
Each & is a complete waveform and the summation must be executed
iamplc:by-sample. An example of this construction is shown in Fig. 8.12
for a motor with 12 slots. If the number of slots/pole is non-integral, the
yoke flux waveform becomes irregular.

8.9 Cogging torque


A further use of the tooth-EMF is the calculation of cogging torque as
pL n
V = it o (8'28)

Where /mag is a fixed current representing the magnet MMF. Determina


tion of fm is left for a future publication, but note that the shape of the
Tcog waveformcan be found without it, if the tooth-EMF flj- is known.

References
Miller TJE, Staton DA and McCilp MI [1993] High-tpeed PC-Based CAD for motor
drives, 5th European Conference on Power Electronics and Applications, EPE9S,
Brighton, 13-16 September 1993.
Ackermann B, Janssen JHH, Soltek R and van Steen RI [1992] New technique, for
reducing cogging torque in a dais of krushJes* DC. motor*, IEE Proceedings 139, No. 4,
315-320.
Boules N [1985] Prediction of no-load flux density distribution in permanent magnet
machine*, IEEE Transactions, Vol. IA-21, No. 4, May/June 1985, pp. 633-643.

8-19
D esig n o f bru sh less perm a n en t -m a g n et m o t o r s

4. Boules N [1984] Two-dimensional analyw of cylindtiad machines with permanent magnet


excitation, IEEE Transactions, Vol. IA-20, No. 5, September/October 1984, pp. 1267-
1277.
5. Hague B [1962] The principle-v of electromagnetum applied to electrical machines -
republication by Dover Publications Inc. of Electromagnetic Problems m Electrical
Engineering (1929J
6. Detnerdash NA, Nehl TW, Fouad FA and Arkadan AA [1984] Analysis of the
magneticfield m rotating armature electronically commutated DC. machines byfinite element*,
IEEE Transactions, Vo. PAS-103, pp. 222^-31.
7. Atallah K, Zhu ZQ and Howe D [1992] Flux waveforms and iron losses fn pervunurti
magnet brushless DC machines, 12th International Workshop on Rare-Earth Magnets
and Their Applications, Canberra, 12-15 July, 1992, 109-120
8. Miller TJE and Rabinovici R [1994] Back-EMF waveforms and core toivs m hntshlcn
DC motors, IEE Proceedings (to be published)
9. Say MG [1947] Performance and design of alternating current machines, Pitman.

&-20
9. CORE LOSSES
9J INTRODUCTION
After copper losses, core losses are generally the second largest
component of power loss or inefficiency in the brushless motor, although
their significance may become overriding at very high speeds. They arise
from the variation of magnetic flux-density throughout the core,
particularly the stator core. This variation incurs hysteresis and eddy-current
losses, (Chapter 16). Briefly summarizing, the hysteresis loss results from
the "unwillingness" of the steel to change its magnetic state, and as the
flux-density varies cyclically the magnetic state describes a locus in the
B/H diagram: the energy loss per cycle is proportional to the enclosed
area, so the average power loss due to hysteresis is proportional to the
frequency of the variation in the magnetic field. Eddy-current loss is also
caused by variations in flux-density, which induce current to flow in the
jtator steel at the same frequency as the variation in the magnetic field.
The EMF which drives these currents is proportional to the peak
magnetic field and to the frequency of variation, but the power loss is
proportional to the square of the EMF and therefore also varies with the
square of the peak flux-density and the frequency. Eddy-current losses
can be reduced by using thinner lamination steels. This is because the
EMF's which drive them are usually in a direction perpendicular to the
plane of the punching. Lamination lengthens the return path for these
currents by forcing them into the circumferential direction, increasing
the resistance. Provided that the frequency is low enough, this causes a
reduction in current and since the power loss is notionally of the form
the reduction in J 2 overwhelms the increase in R and the power loss
is reduced.
In traditional AC machine theory the core loss is viewed as being caused
mainly by the fundamental-frequency variation of the magnetic field at
50 or 60Hz, and this variation is essentially sinusoidal because the supply
is usually a low-impedance utility voltage source. Accordingly the
characterization of core losses in electrical steels has been developed over
the years in terms of Watts per lb or Watts per kg, typically quoted at a
peak flux-density of 1.5T. In brushless motors this fundamental-frequency
lunation is still present, but only in sinewave motors is it even
ipproximately sinusoidal. Most brushless motors, including sinewave
motors, are fed from switched DC sources with pulse-width modulation of
the switches. This means that the applied voltage, and therefore the flux,
9-1
D esig n o f b ru sh less perm a nent -m a g n et m o t o r s
contains many harmonics which may reach frequencies of several tens of
kHz. Although the harmonic components of flux-density are small, they
produce additional core losses over and above the fundamental-frequency
component.
The magnetic field variation in the stator teeth and yoke is primarily due
to the rotation of the magnetized rotor. The time waveshape of the flux-
density in the teeth and yoke is closely related to the space waveshape of
the airgap flux distribution, whose harmonics induce additional core
losses.
A difficulty in the calculation of core losses is that the magnetic flux-
density not only varies in time but also varies widely between different
points in the stator punching. The simplest approach distinguishes two
flux-densities, one for the teeth and one for the yoke, and finer
distinctions are ignored. This makes it possible to obtain manageable
core-loss formulas which include the influence of the main dimensions,
the level of excitation, and the frequency.
Recent attention has been given in the literature to the calculation of
core losses. Bertotti [1] and Slemon [2] recognized the importance of
high-frequency flux pulsations in various parts of the magnetic circuit,
and modified the Steinmetz equation (see below) to accommodate non-
sinusoidal flux waveforms. Bertotti [2] applied the finite-element method
over a range of rotor positions of an induction motor to determine the
flux-density waveforms, subsequendy decomposing these waveforms into
harmonic series and computing the loss components harmonic-by-
harmonic, with various modifications to the material coefficients. Ilowe
[3] reported a similar procedure for the brushless DC motor.
A more direct approach was taken by Slemon [2], who expressed the
basic core-loss equation in terms of the rate of change of flux-density
dB/dt instead of the frequency, and he calculated this rate of change in
the teeth and yoke, assuming idealised waveforms of flux-density in these
sections. A similar procedure has been used in the PCSRD and Pf'cBDC.
computer programs for several years (4,5]. This "waveform" method
produces elegant and simple formulas from which the influence of major
dimensions and parameters can be readily seen.
Most works on core-loss calculation treat the back-EMF calculation and
the core-loss calculation as completely separate exercises, and do noi
9-2
9. C o r e L osses
mention the fact that in a comprehensive design theory the two
calculations must be absolutely consistent with each other. The approach
follow ed here is strictly consistent with the EMF calculation methods
outlined in the previous chapter.

g,2 Nonsinusoidal Steinmetz equation


The specific sinewave loss in W /lb or W /kg is usually expressed in terms
of the Steinmetz equation
^Fe = + Ct B * f2 (91)
which can be rewritten as
__ r dB
Jr >12
(9.2)
^Fe = + dt
The correct value of dB/dt to be substituted is the RMS value over one
complete cycle. 6p is the peak value of the flux-density when the flux is
sinusoidal, and manufacturers loss data is often quoted with = 1.5T.
The exponent n is expressed as a function of Bp since it depends on the
peak flux-density, but often a single value is quoted, typically 1.5-1.7.
The values of the core-loss coefficients Cc, and n can be extracted
from measured core-loss data as described in Chapter 16.

9.3 Core-loss formulas


The specific (i.e. per-lb or per-kg) losses in the teeth and yoke can be
calculated by substituting appropriate values for the peak flux-densities
and their RMS rates-of-change in equation (9.2), and then multiplying
the results by the respective weights of iron in the teeth and yoke.
The calculation of the peak flux-densities and By^ is described in
Chapter 12. The RMS values of dB^/dt and dBy/dt can be obtained from
the waveforms of B j and By as the rotor rotates. Figs. 9.1 and 9.2 show
the ideal trapezoidal forms of these waveforms for the teeth and yoke
respectively, The trapezoidal form arises from the tooth-flux accumula
tion described in Chapter 8. The B j waveform has three different forms
9-3
D esig n o f bru sh less perm a nent -m a g n et m o t o r s

TOOTH
MAGNET

Nl I TS 11.

Tp
r0 - B,
r
0
*

J
(a) w
Fig. 9.1 Toolh flux waveforms for calculation of core Iom
depending on the relative values of the tooth arc aT and the gap
between the magnets With trapezoidal transition shown in Fig
9.1a or 9.1 b ( a ^ s n - PM), the specific eddy-current loss is

WTt 4 A2V Q [W/kg] (9.3)


TT a TT
where /j = 60 x rpm/p is the fundamental electrical frequency in Hz and
p is the number of pole-pairs. For Fig. 9.1c (tx ^ > n - PM), the specific
eddy-current loss is
& 2 -
M) [W/kg]. (9.4)
7T TT

For the stator yoke, with the waveform shown in Fig. 9.2 the eddy-current
loss is
8
[W/kg] (93)
v Pm
where Byp is the peak yoke flux-density.
In reality the tooth- and yoke-flux waveforms are not perfecdy trape-

94
9. C o r e L o sses

Stator Yoke

180
Pig. 9.2 Yolce flux waveforms for calculation of core loss
zoidal, but have rounded corners. For the tooth-flux the effective
transition angle is much closer to the augmented tooth arc (equation
(8.17)) than to the actual tooth arc a T, and is used in equations
(9.3) and (9.4). The augmentation of the tooth arc tends to decrease the
specific eddy-current loss.

9.4 Waveform method


Instead of using the ideal trapezoidal forms of the tooth- and yoke-flux-
density waveforms, the actual waveforms can be used. The flux-density
waveform in the stator teeth is easily obtained from the tooth flux
waveform (Chapter 8):
B .H ) = (9.6)

where Aj. is the cross-section of the tooth (allowing for the stacking
factor of the laminations), and is the angular position of the rotor.
Substituting in equation (8.15),
(9.7)
dt a t

and therefore the RMS value of the j- waveform can be directly used as
it is proportional to the RMS value of dB/dl Equation (9.7) is an
important link between the EMF calculation method and the core-loss
calculation method.
9-5
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

A similar relation is derived for the yoke, since


B ytt) = <f>y (0 (9.8)

so that
<30y 0r(O (9-9)
~dt
and therefore the RMS value of ey can be used in the calculation of the
yoke eddy-current losses. ey is the EMF waveform of a single-turn search
coil wound around the yoke. The RMS value can be extracted from the
waveform using the standard formula for RMS value:

^RMS (9.10)
\
where 5 is the rotor angle in electrical radians. If the waveform eft) or
e(l) is available as a set of N samples e}, covering a completed
period of 2it electrical radians, the integral in equation (9.10) becomes
a sum and
jv
(9.11)

9.5 Augmentation of tooth weight


Where the tooth-flux enters the yoke, there is a transition area in which
it changes direction. In the PC-BDCcomputer program this is accounted
for by augmenting the tooth weight with the weight of the triangular area
shown in Fig. 9.3. The specific core loss calculated in the tooth is
assumed to prevail uniformly through the triangles as well. The theory
underlying this can be seen in the figure. In reality, the transition is
much more complex: see, for example, Fig. 8.6. The weights of the
triangles are not subtracted from the yoke weight when calculating the
yoke losses, and the tooth weight augmentation is applied only when the
waveform method is used to calculate the core loss: it is not used with the
formula method.
9-6
9. C o r e L o sse s

Kig, 9.3 Augmentation of tooth weight with triangle* representing the transition zone
between tooth flux and yoke flux

fig. 9.4 Augmentation of tooth weight viewed over the entire machine crow-section
The weights of the triangles can amount to a 50% increase in the
effective weight of the teeth, Fig. 9.4, and a corresponding increase in
the total computed tooth core-losses. This tends to compensate for any
dif ference between the waveform method in the specific loss calculation.

9-7
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

9.6 Comparison with test data


Ref. [5] gives a limited amount of comparison between the formula
method, the waveform method, and test. However, it should be
recognized that the formulas and methods presented in this chapter are
the simplest ways of estimating the core loss and every opportunity to
correlate such calculations with precise test data should be taken.

References
1. Bcrtotti G, Boglictti A, Chiampi M, Chiarabaglio D, Fiorillo F and Lazzari M [1991]
An improved estimation of iron (tun m electrical machines, IEEE Transaction* on
Magnetics, Vol. 27, No. 6, November 1991, pp. 5007-5009,
2- Slemon GR and Liu X [1990] Core loots in permanent magnet motor*, IEEE
Transactions on Magnetics, Vol. 26, No. 5, September 1990, pp. 1653-1655.
3. Atallah K, Zhu ZQ, and Howe D [1992] The prediction of iron losses in brushless
permanent magnet DC, motors, International Conference on Elcctrical Machines,
ICEM92, Manchester, 15-17 September 1992, pp. 814-818,
4. Miller TJE and McGilp M [1991 ] High-speed PC-based CAD for brushless motor drives,
4th European Conference on Power Electronics and Applications, EPE 91,
Florence, $-6 September 1991, pp. 435-439.
5. Miller TJE and Rabinovici R [1994] Back-EMF waveforms and core losses fn brushless
DC motors, IEE Proceedings 141B, pp. 144-154

9-8
10. ELECTRONIC COMMUTATION
OF SQUAREWAVE MOTORS
10.1 Introduction
This chapter describes the process of electronic commutation in
squarewave brushless DC motors and their controllers. It also describes
the differential equations used by the PC-BDC computer program for the
dynamic simulation of 3-phase squarewave brushless DC motors, with wye,
delta, and unipolar half-bridge connections. The commutation process
is important for detailed understanding of the switching of the power
transistors.
The equations include the freewheeling periods following each
commutation. The integration of the differential equations falls naturally
into "base intervals" of 60 or 120 electrical degrees between successive
commutations. The periodic current waveforms can be constructed from
a knowledge of the currents in one base interval, provided that the
system is in a steady state. The calculation of an isolated base interval
without iteration is possible only if the initial currents can be calculated
algebraically. This is possible at all but the highest speeds, and the
necessary formulas are developed together with the conditions under
which they apply. Many physical insights arise from this analysis into the
structure of the waveforms of the brushless DC motor. Although the
differential equations were developed specially for PC-BDC, which
integrates them using Eulers method, they could equally well be used as
externally-defined system equations in general-purpose simulation
packages.
The importance of a detailed model of the commutation process is
illustrated in the calculation of the no-load speed. Except under special
ideal conditions, this speed cannot always be calculated reliably for
brushless DC motors by means of the simple back-EMF constant used
with DC commutator motors, because the back-EMF waveform may not
be perfectly flat and the supply voltage may be chopped to a lower
effective value than the nominal value. Computer simulation is used to
determine the no-load speed accurately, and the difference between the
no-load speed for motoring and the no-load speed for generating is
discussed.
The quest for computational speed and memory-efficiency places certain
10-1
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

demands on the mathematics of magnetic, electrical, and power-


electronic circuits. In this paper, the circuit equations are developed by
classical methods in a form suitable for efficient computation [1,2,3],
Because of the quest for speed, some generality has been sacrificed. For
example, the number of phases is fixed and the states of transistors in
the controller are restricted to those which occur during normal
operation. Although the analysis could be extended to include fault
conditions, this is not included here. Matrices are not used: instead, the
differential equations for every mesh in the circuit are written individu
ally and manipulated into a form suitable for efficient computation, and
detailed formulas are developed for the initial and expected final
currents at the ends of each base interval. In these respects the analysis
differs from earlier work by Dcmerdash [4] and Evans [5].
The physical interpretation of the equations is especially useful in
connection with the power electronic controller. Most introductory works
on the brushless DC motor ignore the details of the process of
commutation, yet this is crucial in the sizing of power electronic devices
(transistors and diodes) and in determining the limits of operation (for
example, the maximum attainable speed).

10.2 Basic Principles


The circuit diagram is shown in Fig. 10.1 for the wye-connected motor
and Fig. 10.2 for the delta-connected motor. The sequence of switching
transistors is shown in the corresponding waveform diagrams, Fig. 10.3
(wye) and Fig. 10.4 (delta). The waveforms in Figs. 10.3 and 10.4 are
idealised: they assume that the generated EMF waveforms are trapezoidal
with flat tops of sufficient width to produce constant torque when the
line currents are perfectly rectangular 120 waves. (This idealization is
only in the diagrams and not in the equations). The only assumptions
are that the generated EMF waveforms are known, and that the self- and
mutual inductances are constant (i.e. independent of both current and
rotor position).
Under these conditions the instantaneous airgap torque is given by
T = J [e ,/, + Nm, (10.1)

10-2
10. E l e c t r o n ic C o m m u t a t io n o f S q u a r ew a v e M otors

Fig- 10.1 Circuit diagram of wye-connected motor

Fig. 10.2 Circuit diagram of delta-connected motor

For both wye and delta connections, Figs. 10.3 and 10.4 reveal that the
waveforms repeat every 60 electrical degrees, with each 60 segment
being "commutated" to another phase. For example, in Fig. 10.3 the

10-3
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

segment of line current A between 30 and 90 "reappears" as - ^ i n the


next segment from 90 to 150, and then as 4-^ in the following segment
from 150-210. The "route" followed by such a segment can be
characterized by the sequence of signed suffixes associated with the lines
(or phases). For the example segment just quoted, the sequence is
1,-3,2,-1,3,-2,1,.. This is easy to remember as the negative phase
sequence 3,2,1,3,2,1... but with every other sign reversed.
The commutation of the 60 segments means that the entire set of line
and phase waveforms can be constructed from one 60 segment when
the motor is operating in the steady state at constant speed.
In order to calculate the currents in one segment it is necessary to
integrate the voltage equations through the 60 base interval. To do this
we must first formulate the differential equations, then determine their
initial conditions, then integrate them. Unfortunately there is no natural
point on the waveforms where all the currents are zero, so the initial
values at the start of the base interval must be determined as shown
below. An alternative procedure is to start with zero currents and
integrate several consecutive periods until a steady state is reached.
However, this takes a long time and generates much waveform data that
is wasted. Therefore, we will tiy to calculate the initial conditions a priori
The actual voltages Uj, v3 across phases 1,2,3 are expressed by

= e{ + J?i\ + Lpx + Mp2 + Mp^ (1 0 -2 )

k2 = ^ + Mpl + Lp^ + Mpi ( 1 0 .3 )

v3 = % + Riz + Mp 1 + Mpz + Lpj (10.4)


where pl = di^/dt, etc., and ij, i^, and tj are the phase currents. These
equations are independent of the external connections, whether wye,
delta, or unipolar half-bridge.

10-4
10. E l e c t r o n ic C o m m u t a t io n o f squarew ave M otors

-30 0 30 60 90 120 150 180 210 240 270 300 330 0 lec
'SP Lin<3

I\
'A Cur rents
/
y

\
f
'B

'C
\
Ph ase
ENIFs

/
e1
' /
>

f
e2

1
Unis-Lin e
EM
e12 1
<D
CM

= e 1~

Q5 Q1 Q3 Q5
Q6 I Q2 I Q4

Fig. 10,3 Waveforms of wye-connected motor

10-5
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

-30 0 30 60 90 120150 180 210 240


1 270
I I300 330 lec
'a I IspL---
s p 1------ lin e'
LINE
= i1 -- ii.3 CURRENTS
CURRENTS
/
.
'b
T
______ _
' c ___________________________________________

-----------------------
PHASE
PHASE
61 EMF
EMFxv
------------------------------------------------
X
2/3*1
2/3 * leo
SP
1 / 3 * ls
spp / ,SP -----------------------------

A\. ----- ------------------------ PHASE


PHASE -
*1------------------------------------------------------ -CURRENTS
CURRENTS

.3 ------ iJ.-----
f f

Q5 Q1 Q3 Qi
06 02 Q4

Fig. 10.4 Waveforms of delta-connected motor

10-6
10. E lectronic Com mutation o f Squarewave M otors
10.3 Circuit Equations - Wye
10.3.1 Commutation
Commutation is initiated by a switching event and is characterized by the
currents in two separate meshes. Following the switching event, one of
these currents builds up from zero and the other one decays towards
zero. The switching event taken as the initiator of the "base" interval is
the turn-off of Q5 and the simultaneous tum-on of Q l. In Fig. 10.3 this
interval is from 30 to 90. The two mesh currents are as shown in the
top-left diagrams of Figs. 10.5 and 10.6 respectively. Fig. 10.5 shows the
current building up through Ql and Q6 in phases 1 and 2 (lines A and
B). Fig. 10.6 shows the current freew heeling through diode D2 and
transistor Q6 in phases 3 and 2 (lines C and B). Ideally the (negative)
current in line B should remain constant, while the line C current falls
to zero and the line A current builds up to the set-point value.
(a) Building (b) Q1 chopped

(c) Q6 chopped (d) Over-running

10-7
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a c n e t m o t o r s
(a) Freewheeling (b) Q1 chopped

(c) Q 6 chopped (d) Over-running

Fig. 10.6 "Freewheel" mesh during commutation and chopping - wye connection

The currents in the corresponding diagrams a,b,..d in Figs. 10.5 and 10.6
are flowing simultaneously. The reason for having two separate diagrams
is to associate them clearly with their respective mesh-voltage equations.
It is assumed, initially, that the set-point current /sp is flowing through
phases 3 and 2 in series, so that ig = / and ^ = - / . All branches of the
circuit are inductive, so that when Q5 switches off the current in line G
continues to flow, causing diode D2 to become forward-biassed, clamping
the positive terminal of phase 3 to the negative rail. At the same time,
although Q l is on, the current in line A is initially zero. The positive
terminal of phase 1 is held at +V^. The potential of the negative terminal
of phase 1 is not known, but it is normally at a potential lower than +VS,
so the current ij begins to rise. In a similar way, the voltage across phase
3 is normally negative and therefore begins to decay.
10-8
10. E l e c t r o n ic C o m m u t a t io n o f Sq u a r e w a v e M otors
'The process continues until ^ = 0 , at which point it remains zero
because D2 switches off and there is no conduction path in line C. There
is then only one mesh for current to flow, i.e., the loop through phases
1 and 2. This current continues to rise until it is limited by the chopping
action of the current-regulator, which turns transistor Q1 (or Q6) on/off
to maintain the current at the set-point value /Jp. If there is insufficient
voltage for the current to reach 7 , then it is limited instead by the
resistances, inductances, and back-EMFs of phases 1 and 2 in series. This
condition tends to arise at high speed.
10,3.2 Period A and Period B
Period A is the initial freewheeling period, just after switching, when
both mesh currents are non-zero. During Period A, all three lines are
conducting. The subsequent period when only phases 1 and 2 are
conducting is called Period B. Period B begins when the freewheeling
mesh current extinguishes. During Period A, ^ is decaying and i, is
increasing. Ideally di^/dt = - di^/dt so that ^ remains constant. It is
shown in Fig. 10.7, which also shows the reverse recovery of D2. Any
inductance in line C (including motor inductance) multiplies the di/dt
of the reverse-recovery current in D2 to produce a positive voltage spike
across Q5 (not analyzed here).

PERIOD A
PERIOD B

Isp

Q5 turns off Reverse recovery of D2


Q1 turns on

Fiji 10.7 Expanded current waveforms during commutation

10-9
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

10.3.3 Chopping (regulation)


When the current in line A builds up to the set-point value /spl either Ql
or Q6 may be turned off by the action of a current-regulating circuit If
the control strategy is voltage PWM rather than current-regulation, either
of these transistors may be turned off before iA reaches / . The effect of
turning off Ql on the two mesh currents is shown in Figs. 10.56 and
10.6& Similarly the effect of turning off Q6 is shown in Figs. 10.5c and
10.6c. Voltage-mesh equations can be written for all of these conditions.
10.3.4 State-space averaged voltages
The voltage ^bld controlling the build-up of current in phases 1 and 2 is
the line-line voltage across these two phases in series, which can be
written
^ b ld = v n = *1 ~ vi v a b - (10.5)
Similarly the voltage ("fwh" = "freewheel") controlling the decay of
current in phases 3 and 2 is the line-line voltage across these two phases
in series, which can be written
( 10.6)

If Ql is the controlling (chopping) transistor, the value of Vbld is


determined by the state of Q l, which is labelled Jqj :
Vs - 2 V q - 2 R qil (sQl=l) (10.7)
bW " ~yq - V i (^Qr)

or

The value of ^fwh is not affected by the state of Ql:


*fwh = K i+ v * qK ) - (10.9)

10-10
10. E l e c t r o n ic C o m m u t a t io n o f S q u a r ew a v e M otors
If Ql is chopped at a very high frequency then Sqj can be set equal to
the duty-cycle (mark/period ratio) of Ql, and in this case yb)d can be
interpreted as the average voltage across the terminals AB. (This is the
principle of state-space averaging).
The state-space average value is useful in algebraic calculations of the
steady-state currents; in this case Sq j will have a fractional real value. But
the same formulas for ^bld and can be used in digital simulation,
where Jqj is the actual state of the chopping switch; in this case Jqj has
a binary value (on or off).
Similar equations result if Q6 is chopped instead of Ql: the result is

Hu =W 2V 2Vi>+ W i - H). <10-10)


which is identical to equation 5 except that Jqj isreplaced by w .
However, when Q6 is chopped it has an effect on the freewheeling mesh:

^wh = + V i + Kd>+ a - + 2 ^ )- (1 0 n )
When the line A current is building up there is no difference between
chopping Ql and chopping Q6. However, chopping Q6 will accelerate
the decay of the freewheeling current in line C by connecting the supply
voltage against the flow of freewheeling current during the periods when
Q6 is off.
10.3.5 Euler form of voltage equations (wye connection)
In the wye connection
h + h. 4 h ~ 0 (10.12)
and therefore
Px* P i* Pi = 0. (10.13)

If equation 13 is substituted into equations 2-4, we get


vx - Cj + Ri\ + (L - M)px (10.14)

10-11
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

gj + + (L - A/)A (10.15)
(10.16)
It is now a matter of algebra to solve equations (10.14-16) for the
derivatives fa, p%, and py the result is
(10.17)
3 L>
~ J'fwh + 13R ~ L 'P \ (10.18)
1 2L
Pi = - Pi ~Pi (10.19)
where
enx ~ + + h) ( 10.20)

~ % ~ ~ i + 2yj)
r ( 10 .21 )

L = L - M, ( 10.22)

Equations (10.17-19) are in the form required for integration, e.g. by


Eulers method or the method of Runge-Kutta. Actually equation (10.19)
is not needed, because once ij and ig are determined by integrating
equations (10.17) and (10.18), ^ can be determined from equation
(10.12) directly.
As a check on the correctness of these equations, and also to gain
additional insight into the working of the circuit, consider the special
case that arises if Vj, V^, R^, and J?are all zero, while Cj = and = 0.
Then, if Q1 is the chopping transistor, the equations for pi and ps
reduce to
(10.23)
and
(10.24)
which shows that if there were no generated EMF then 2/3 of the supply
voltage (modulated by Jq ,) would be dropped across phase 1, and 1/3
10-12
10. E l e c t r o n ic C o m m u t a t io n o f S q u a r ew a v e M otors
across the parallel combination of phases 2 and 3 during the freewheel
ing period.
In Period B there is only one mesh and one current (as in Fig 10.5),
and iy = -i%, with *3 = 0. The solution for pj is
Pi = - J i [ Ku - ck - e,) - 2R/\]. (10.25)

10.3.6 Initial conditions (wye)


The initial conditions for the 60 base interval are the same as the final
conditions for the previous 60 interval but "commutated" according to
die sequence defined earlier. Having written the circuit equations for the
base interval, it is easier to find the final values for this interval and then
use the commutation sequence to transfer them to the beginning of this
interval. If the subscript F denotes the final value at the end of the base
interval and S denotes the starting value, then the re assignment will be
according to
'is " hv (10.26)
h s ~ - / 1F (10.27)
(10.28)
I
1
V3

To find the final values ijF, ^ F, and ^ F, assume that by the end of Period
B the line current iy = has arrived in the hysteresis-band of the
current-regulator, and that this band is a very small percentage of /s ; see
Fig. 10.7. The current tj is essentially a DC value. rHiis is possible only if
the duty-cycle of the chopping transistor Ql is less than 1. From equation
(10.25), with py = 0 (steady-state DC),
. Hid ~ (ei ~ 5 ) (10.29)
1 2R
If we set iy = I (the set-point value) then the duty-cycle of Q l can be
calculated from equations (10.8) and (10.29) as
= gi - H * 2 RIsV + VA * \ + ^ / s p (10.30)
K ~ K + V, - Rq I
10-13
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

If this expression gives Jqj < 1, then the initial conditions can be
assigned from
7IS = 0 (10.31)
4s 4p (10.52)
^3S = V (10.33)
If, on the other hand,turns out greater than 1, it means that there
is insufficient voltage to drive the current up to the set-point value of the
regulator. This can mean one of two things. Either the set-point current
is greater than the resistance-limited DC value of the current at the end
of the 60 interval. Or, the current is never reaching a steady state within
the 60 interval. In the first case the final steady-state DC value can be
calculated as
J \r = Vs ~ 2 *q - (C[ " (10.34)
,F 2(R + R J
with = 1, (The current-regulator will be saturated in this case and Ql
will remain on for the entire 60 interval). The second case can arise at
high speed. Unfortunately there is no way thefinal current can be calculated
analytically. What is done instead is to run a trial 60 integration with the
initial conditions calculated by equations (10.31-33) or (10.34), and then
test the final values, suitably "commutated", to see if they match the
starting values. If they do not, then the final values from the integration
are "commutated" into the starting values and a second 60 integration
is run. The procedure is repeated until the final values and starting
values converge to within a predefined tolerance.
When Q6 is the chopping transistor, the equations are identical except
that .?Qt is replaced by Jq 6 in equation (10.30).

10.4 Circuit equations - delta


10.4.1 Commutation
Fig. 10.8 shows the condition of the circuit immediately before the
instant when Q5 switches off and Ql switches on. It is assumed, initially,

10-14
10. E l e c t r o n ic C o m m u t a t io n o f S q u a r ew a v e M otors

** Is p /3
Fig. 10.8 Condition of ciclta-conncctcd circuit jujt before commutation of Q5/Q1
that the current 7sp is flowing in line C: that is, V through phases 3
and 1 in series, and 2/,p/3 in the reverse direction through phase 2, so
3
that i, = i, = 7sp/3 and i> = -2/jp/3.
The 2:1 ratio between the currents in the two parallel branches of the
delta is intuitively clear if it assumed that the currents are resistance-
limited while the sum of the EMFs e-y+e^ = -e^. This EMF condition is
satisfied by the EMF waveforms in Fig. 10.4, but it is not always the case
because it depends on having the correct magnet pole-arc and coil pitch.
(See Chapter 5). When the current is being regulated (chopped) by Q1
to a value within the hysteresis-band around J , the 2:1 division may not
be obvious, but is proved below.
AU branches of the circuit are inductive, so that when Q5 switches off the
current in line C continues to flow, causing diode D2 to become forward-
biassed, exactly as in the wye circuit This clamps the positive terminal of
phase 3 close to the negative rail, as shown in the freewheeling circuit
diagrams of Fig. 10.10. At the same time, although Q1 is on, the current
in line A is initially zero. The positive terminal of phase 1 is held at
+Vt~Vd. The negative terminal of phase 1 is held close to zero by Q6, so
that normally the current tj begins to rise. The circuit condition is shown

10-15
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

in Fig. 10.9a. Note that there is as yet no current in line A, even though
Ql is on.
(a) Main conduction mode (b) Q1 chopped

{c) Q 6 chopped (d) Over-running

Fig. 10.9 "Build" meih during commutation and chopping - delta connection
By similar reasoning the voltage across phase 2 is close to zero, and the
current in phase 2 therefore freewheels. The voltage across phase 3 is
~(VS-Vd) initially, and the current jg begins to decay. The freewheeling
mesh is shown in Fig. 10.10a. This process continues until Iq reaches
zero, at which point D2 switches off and begins to increase in the
opposite (negative) direction. There is now no conduction path in line
C. There is only one loop for current to flow through the converter, that
is, the loop through Ql and Q6, assuming that both transistors remain
on. The current in this loop divides between the two parallel branches
in the delta-connected motor, and eventually it is expected that a steady-
state will arise with ij = 2/sp/B and ig = ^ = - /sp/3, before the next
commutation. This final condition is shown in Fig. 10.9a.

10-16
10. E l e c t r o n ic C o m m u t a t io n o f S q u a r ew a v e M otors

(a) Freewheeling (b) Q1 chopped

e f hjri 4 >h
A
\ 4

l A ! > ) ' \ }

(c) Q 6 chopped (d) Over-running

10.4.2 Period A and Period B


The initial period when all three lines are conducting is called Period A.
The subsequent period when only lines A and B are conducting is called
Period B. During Period A, di^/dt is negative and *A is increasing. Ideally
di^/dt = -diof dt so that ig remains constant.
At the same time, it is desirable that di^/dt = -di^/dt and that remain
constant during the same period. Period A is the freewheeling period. It is
too rapid to appear on Fig. 10.4 but is shown in expanded form in Fig.
10-7, as before.

10-17
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

10.4.3 Chopping (regulation)


As in the wye-connected motor, the current in line A rises until it is
limited by the chopping action of the current-regulator, which turns
transistor Q l (or Q6) on/off to maintain the current at the set-point
value I , assuming that it ever reaches this level. If there is insufficient
voltage for the current to reach /sp, then it islimited instead by the
combined resistances, inductances, and generated EMFs between
terminals A and B of the delta. This condition tends to arise at high
speed.
10.4.4 State-space averaged voltages
The same "build" and "freewheel" voltages appear at the terminals of the
converter as before. Again, Jqj is either the state-space averaged duty-
cycle of Q l, or its binary state, and likewise for The "build" and
"freewheel" voltages are defined by the connection equations
Hid = v\ = ~(yi * Ks) (10.35)
Hwh = v2 = "(*i + <10-36)
where Vj, are the voltages across phases 1,2,3 expressed by
equations (10.2-4).
When Ql is the controlled (chopping) transistor, from Figs. 10.9a and
10.96,
Hid = sQl[ Vs - 2 Vq - J?q(2,* -4 -4 )] 1097J
* (1 - W

and from Figs. 10.10a and 10.106,


Hwh = V H +4,(4 h)- (1038>
When Q6 is the controlled transistor, from Figs. 10.9a and 10.9c,
Hid = *Qst H " ^ Vq ~ n o 39)
+ (1 - ^ ) [ - H - Vq - ^ (/;-/-)];

10-18
10. E le c t r o n i c C om m utation o f Squarew ave M o to r s

and from Figs. 10.10a and 10.10c,


*fwh = V + + *q(V 4)] + d " V I +2^h (104)
Equations 37-40 are similar to equations 8-11 for the wye connection.
10.4.5 Euler form of voltage equations (delta connection)
The Euler form o f the differential equations is given by the solution for
and in equations (10.41-43), with equations (10.44-49):

Pi (10.41)
L" -
M"

Pi - ~ M '* * Rn ] (10.42)

Pi = j l vo + RqO] - h ) ~ R h ~ M (px * A )] (10.43)


where

I" (10.44)
II

M" (10.45)
*
U

R\i ~ R]\ ~ ~ ~ h (10.46)

R*i ~ ^ - & (10.47)

*fwh - l) (10.48)
^ L (

(10.49)
Hig;
1
Jb
i

i
D.

z, fwh

10-19
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

In Period B there is only one active mesh, but this divides into two
parallel branches in the delta-connected motor. Accordingly there are
two differential equations. With D2 off,
4 = h, (10.50)
VUA = v\ = ei + R i\ + L P \ + 2 M Pi< (10.51)
and
= ~r2 ~v3
Kbld (10.52)
= - ( V ^ ) - ZOh+Jb) -
Solving these equations for p\ and p%, we get

ks7 - *4 *
Pi =
81 1 l +M 2
(10.53)
, _ 2 A/2 1
L* M\
_ ["( ^bld + " 2 A//7t|
A 2(+A0
where

10.4.6 Initial conditions and final DC values (delta)


As in the case of wye connection, the initial conditions for the 60 base
interval are the same as the final conditions for the previous 60 interval,
but "commutated" according to the sequence defined earlier. Having
written the circuit equations for the base interval, it is easier to find the
final values for this interval and then use the commutation sequence to
transfer them to the beginning. If the subscript F denotes the final value
at the end of the base interval and S denotes the starting value then the
re-assignment will again be according to

10-20
10. E l e c t r o n ic C o m m u t a t io n o f Sq u a r e w a v e M otors
(10.57)
(10.58)
To find the final values ijF, and i^y, again assume that by the end of
Period B the line current has arrived in the hysteresis-band of the
current-regulator, and that this band is a very small percentage of isp; see
Fig. 10.7. The current A is essentially a DC value, and can be regarded
as a DC source fed to the two parallel circuits of Fig. 10.11, from which
the following equations can be written down:
Hid ~ e i + R i A (10.59)
Hid = + **j) - ( 4 + (10.60)
Note that i,, ig and i$ have all been taken as positive when flowing away
from the line terminals A,B, and C respectively, as in the original circuit
diagram, Fig. 10.2. The line current iA is given by
ja = h h- (10.61)
Solving these equations for i1F and igF,

/IF - 2 ; -
3 A
+ % * *
3 /?
(10.62)

(10.63)

O o

PHASE 1

'2 = i3 I PHASE 3 PHASE 2


\< I +H^
is'3 '2b

Rg- 10.11 DC conditions in delta circuit


10-21
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

In an ideal brushless motor designed for delta connection,


e, + Cj + e, = 0 (10.64)
and therefore
- |v (10.65)

while
% - j ' . - \ h v <1 )
The current in the branch containing only one phase is thus twice the
current in the branch containing two phases. Since equations (10.62-66)
do not contain Vb1d> they are independent of the state of the chopping
transistor Q l and are valid for all values of duty-cycle including 1.
Equations (10.62) and (10.63) show that if the loop EMF e, + ^ + % is
not zero, the division of current between the phases of the delta is
disturbed from the ideal 2 /3 :l/3 ratio. In a severe case this can prevent
the motor from reaching the desired torque, and it may add considerably
to the losses.
Of course, the attainment of a steady DC line current at the end of
Period B is possible only if there is sufficient voltage, i.e., if Jqj s 1. For
a given value of iA (e.g., iK = /sp), the duty-cycle, Vqj can be calculated
directly from equations (10.37),(10.59) and (10.65) as
_ el * (2R/3 + # q)^.p Vd + Vq (10.67)
K - K, + t'd ^
If this expression gives Jqj < 1, the initial conditions can be assigned
from equations (10.56-58), with *j, ^ and ^ obtained from equations
(10.62-63). If, on the other hand, Sqj turns out greater than 1, it means
that there is insufficient voltage to arive the line current up to the set-
point value I . This can mean one of two things. Either the set-point is
greater than the resistance-limited DC value of the current at the end of
the 60 base interval. Or, the current is never reaching a steady state
within the 60 interval. In the first case the final steady-state DC value
can be calculated from the DC loop equations with ^ = ig :

10-22
10. E l e c t r o n i c C o m m u ta tio n o f S q u a re w a v e M o t o r s

* 2V 2 *qO\ ~ h) = + R l\ (10.68)
= - ( e j + ,) - 7Ri3
r j 3 + /?y,
2 + (e , + Cj + = o (10.69)
to which the solution is
(10.70)

J2F = J3F = (10.71)


with Jqj = 1. (The current-regulator will be saturated in this case and Q]
will remain on for the entire 60 interval).
The second case (failure to reach a steady-state current) arises at higher
speeds and, as in the case of wye connection, there is no way the final
c u rre n t can be calculated analytically. What is done instead is to run a
trial 60 integration with the initial conditions calculated by equations
(10.56-58) with (10.62-63), and then test the final values, suitably
"commutated", to see if they match the starting values. If they do not,
th en the final values just calculated are "commutated" into the starting
values and a second 60 integration is run. The procedure is continued
until the final values and starting values converge to within a predefined
tolerance.
W hen Q 6 is the c h o p p in g transistor, the equations are identical except
that Jqj is replaced by Sq^ in equation (10.67).

10.5 Unipolar half-bridge 3-Phase controller


10.5.1 Commutation
fig. 10.12 shows the circuit diagram of this 3-phase system. The phases
are electrically independent, though they are still coupled by mutual

10-23
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

Building Freewheeling
Current Current

Fig. 10.12 Half-bridge unipolar 3-phase circuit diagram


inductance. With only three transistors, the phase currents are
unidirectional. The ideal waveforms shown in Fig. 10.13 are 120
rectangular waves with all the negative half-cycles missing. 120
conduction makes it possible to produce torque at all rotor positions,
provided that the phase EMFs have sufficiently wide flat tops.
The commutation sequence is based on a 120 base interval, but in other
respects it is similar to that of the wye and delta cases; indeed it is
simpler because at least one of the phase currents is zero at any instant
10.5.2 Period A and Period B
The freewheeling and "build" meshes are highlighted in Fig. 10.12. The
base interval has the same starting point as for the wye and delta cases,
that is, when Q5 turns off and Ql simultaneously turns on. decays by
freewheeling through D2 and the damping or "suppression" resistor
while ij builds via Ql and the supply Vs. Meanwhile, ^ = 0. The
freewheeling and "build" meshes are electrically separate.
10.5.3 Chopping (regulation)
The current in phase 1 rises until it is limited by the chopping action of
the current-regulator, which turns transistor Ql on/off to maintain the

10-24
10. E lectronic Commutation o f Squarewave m o t o r s

120 elec deg


Period Period
B Ideal rectangular current waveform

' 7 T
Freewheeling
current

e1 , Phase 1 EMF

J
| 30 90 150
0 180
_ .... Current regulated
current^ y chopping

Actual current waveform

Freewheeling current
(May contain chopping ripple
coupled from other phases)
Fig. 10.13 Half-bridgc unipolar 3-phase waveforms
current at the set-point value 7jp) assuming that it ever reaches this level.
I f there is insufficient voltage for the current to reach /5p, then it is
limited instead by the combined effects of resistance, inductance, and
EMF. This condition tends to arise at high speed.
10.5.4 State-space averaged voltages
The "build" and freewheel voltages are defined by the connection
equations
Kbld = (1 0.72)

10-25
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

- v2. (10.73)
where v,, v^, are the voltages across phases 1,2,3 expressed by
equations (10.2-4).
Ql is the controlled (chopping) transistor, and
* *Q.[ k - yq - v ) + (i - v [- - v j; (10-74>

Kfwh = - <10-75)

10.5.5 Euler form of voltage equations (unipolar 3-phase connection)


When equations (10.72-75), (10.2-4) are solved for p^ and p$ the result

* = L -1 M,.2 [L 4 - M yl ] (10765

* =ri
L -. *M1~MV'+ L V' ] * (10>77)
where
<= - R h\ (10-78)
r ' = - Vd - (JR * - e, (10.79)

These equations hold until i^ reaches zero. Then the freewheeling period
A terminates. In the following period B only phase 1 conducts. The
controlling equation is:
/>. = 3 <10-80)

10.5.6 Initial conditions and final DC values


The initial value for is equal to the final value 1F (which is the same
for all phases). The resistance and EMF-limited value of i1F is

10-26
10. E l e c t r o n ic C o m m u t a t io n o f S q u a r ew a v e M otors

(10.81)

For a given set-point current / , the required duty-cycle can be worked


out along the same lines as before:

(10.82)
*Q I = vs- V H +( 4 r 4 , H p

Jqj cannot take a value greater than 1. If equation (10.82) gives a value
greater than 1, the regulator will be saturated with Jqj = 1, and i1F will
be given by equation (10.81) with Jqj = 1:

10.6 Over-running
Another conduction mode which occurs in all of the circuit configura
tions is the regeneration mode, when the generated line-line EMF
exceeds the supply voltage. In this case the current may flow in the
reverse direction back to the supply, as shown in Figs. 10.5d and 10.6d for
the wye connection and Figs. 10.9d and lO.lOd for the delta connection.
Referring to Fig. 10.5 if, reverse current flows through the diodes D1 and
D6, and these diodes communicate the supply voltage to the series
combination of phases 1 and 2 in opposition to the line-line EMF. The
reverse current is unaffected by the chopping of either transistor Q l or
because it flows through D1 and D6. Accordingly, Fig, 10.6d shows
Only the freewheeling current in phases 2 and 3, which is independent
of the reverse current in phases 1 and 2.
This condition tends to arise at speeds close to, or above, the no-load
Hid - K - 2 vd (10.84)
speed, when the load may be overhauling or over-running the motor.
The voltage equations are as follows : for both the wye and the delta
connections, where Vbld is related to the motor voltages by equations
(10.5) and (10.35) respectively.
10-27
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

The simplest example of over-running is the plain generating condition,


in which the transistors are never switched on. In this case the line
current begins to flow in the reverse direction as soon as the generated
EMF component of the motor voltage exceeds the supply voltage (see
equations (10.84,10.5,10.35, and 10.1-4). Conduction through the diodes
continues until a natural point of commutation, and the converter circuit
acts as an uncontrolled three-phase bridge rectifier. This plain generating
mode is unusual in squarewave brushless machines: normally, PM
generators are designed for sinusoidal EMF, However, the same
principles of commutation apply in both cases.
The over-running condition is more likely to be met in squarewave
motors running at or just above the no-load speed, where it exists for
only a fraction of the cycle. This part-cycle over-running may be inherent
in the operation at no-load, and accurate calculation of no-load speed
may require a dynamic simulation. This is discussed in the next section.

10.7 Practical examples and comparison with test data


10.7.1 Comparison of measured and computed waveforms
An exacting test of the simulation equations is the evaluation of the line
current waveform under high-frequency chopping, because of the large
number of commutations. With vollage-PWM the test is more demanding
than with current regulation, because there are no fixed limits between
which the current is chopped. An example of such a test is shown in Fig.
10.14, which shows calculated and measured waveforms for a small wye-
connected brushless DC motor obtained with the PC-BDC computer
program. The control strategy is voltage-PWM, meaning that the
controlling transistor is chopped at a fixed frequency but with a variable
duty-cycle (variable off-time) in response to the torque-demand signal.
Agreement between the test data and the calculated waveform is self-
evident. Clearly the calculation depends on having an accurate calculated
waveform of the back-EMF, and although PC-BDC has the facility for
using external or measured back-EMF waveforms, the EMF waveform
used to calculate Fig. 10.14 was calculated internally by the program
using the tooth-flux accumulation method reported in Chapter 8.

10-28
10. E l e c t r o n ic C o m m u t a t io n o f S q u a r ew a v e M otors

PHASE CURRENT UflUEFORn


PHASE CURRENT <flflPS> x l.OeO
1.401
xx TEST
1.201
1.00-
0. 80-

0 . 60 -

0. 40-

0 . 20 -

0 .4 0 0. BO 1 .2 0 1 .6 0 2 .0 0
ELECTRICAL DEGREES x 1.0 e 2

Fig. 10.14 Comparison of tested and calcu lated curT en t waveforms

Normal commutation zone

10-29
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

10.7.2 Accurate calculation of no-load speed


The no-load speed jVq is often calculated from the simple classical DC
motor formula
AJj = ~ x rev/min, (10.85)
% 2w
where is the DC supply voltage and kE is the back-EMF constant (see
Chapters 1,5). This equation ignores the effects of resistance, transistor
volt-drop, and friction and windage. It works properly only if the line-line
back-EMF during the normal conduction period or "commutation zone"
is perfectly flat-topped, with a value that equals the supply
voltage Vs at the no-load speed. This ideal condition is shown in Fig,
10.15.
In practice the line-line back-EMF waveform is not perfectly flat-topped
throughout the conduction period, as in Fig. 10.16. There is now no
simple relationship between Va and e ,, at the no-load speed.
The no-load speed is defined as the speed at which the shaft torque is
zero, when the motor is fed with correctly commutated voltage pulses of
the rated value. Ideally the voltage pulses should not be chopped by any
current-limiting or regulating circuitry if this definition is to be
meaningful. At the no-load speed, the motor generates just enough
torque to overcome the windage and friction loss, while the DC supply
supplies power exactly equal to the combined losses in the motor and
controller.
Normal commutation zone

10-30
10. E l e c t r o n ic C o m m u t a t io n o f S q u a r ew a v e M otors
This condition can be simulated on the computer, but it cannot be
calculated directly (unless the back-EMF is perfectly flat and all volt-drops
ar.d losses are known exactly).
Al the no-load speed, it is possible for the back-EMF ^ to exceed the
supply voltage for short periods during the commutation zone, as shown
in Fig- 10.16. When this happens, di/dt becomes negative and the line
current tends to decrease; it may even become negative as shown in Fig.
10.17. The simulated current waveform shown in Fig. 10.17 is for a 4-
po)c. 12-slot machine at 3750 rpm, with windage and friction loss
arbitrarily set to zero. The electromagnetic torque waveform has an
average value of practically zero, and since the windage and friction loss
U set to zero, this is also equal to the shaft torque. The no-load speed is
therefore 3750 rpm, compared with the value 3655 rpm calculated from
equation (10.85), a difference of 2.5%.
The windage and friction power of the example motor is proportional to
speed cubed and is 7W at 3000 rpm. When these figures are incorpo
rated in the computer simulation, the no-load speed falls to 3625 rpm.
The corresponding waveforms are shown in Fig. 10.18. This value is
closer to the value calculated from equation (10.85), but the apparent
l.. i -t. (y a w oMtHii,
Tftapa] * .0*0
0.BQ

[\ f\ \ y - \\
-am
CUoltrl 1.0*1
<.00 --------a ^ x c w d * [24V] atthla point

-------x y -----
^ ____ ^ ^
-2.00
-4.00
rM-i K 1.0t-2
tf.oa /------ -M e a n torque * 0
4.00
2,Du

-a.tfr
/\ / T a A A A S.KK^.SO s.
-4.00
__ Doip*- BOfiu#n [l*e dol x 1.0*2

fig- 10.17 Line current (top), line-line EMF (centre), and torque waveforms at 3750
rev/min, with windage and friction assumed zero
10-31
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

Fig. 10,18 Line currcnt (top), line-lincEMF (centre), and torque waveforms at 3625
rev/min, with windage and friction equal to 7W at 3000 rev/min

accuracy of the simple formula is fortuitous, because the error due to the
neglect of windage and friction in the simple formula is compensated by
the error due to the assumption of a perfectly flat-topped back-EMF
waveform. The true no-load speed must be determined iteratively by
successive simulations. At no-load the difference between Vt and e11 is
just enough to raise the current to the level needed to equal die windage
and friction torque.
An exacting test of the no-load speed calculation is to compare it with
the measured value, and the results of such a test are summarised in
Table 1 for the motor whose current waveform is shown in Fig. 10.14. In
this case the simple formula is unable to predict the correct value, even
when the supply voltage is multiplied by the modulating index or duty-
cycle of 0.30. The simulated waveforms for this condition are shown in
Fig. 10.19.

10-32
10. E l e c t r o n ic C o m m u t a t io n o f S q u a r ew a v e M otors

No-load speed (rev/min)


M easu red a t Vs - 167 V, 3476
IhUy-CycU o f voltage-PW M = 0.30
C alcu lated fro m V/Aj. 7526
C alcu lated fro m 0.30 x Vt / k ^ 2258
S p e ed a t w hich sim u latio n shows zero 3472
lh a ft to rq u e

T able I
M easured and calculated N o -L oad S peed

Fig. 10.19 Simulated line current (top), line-line EMF (centre), and torque
waveforms for motor of Table 1 at 3472 rev/min
Further insight can be obtained by considering what happens as the
motor speed rises steadily, with fixed commutation angles, fixed supply
voltage, and no current limit. At low speeds and the current
Waveform is generally positive, with positive torque much greater than the
windage and friction torque. At the no-load speed, the electromagnetic

10-33
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

torque isjust sufficient to overcome the windage and friction torque. The
DC supply current is still positive, as it supplies all the losses.
At higher speeds, the shaft torque becomes negative, and a mechanical
prime mover is required. At first, at speeds only slightly higher than the
no-load speed, the electromagnetic torque may still be positive
(motoring), even though the shaft torque is negative due to the windage
and friction. At some speed the electromagnetic torque is zero: the prime
mover is then supplying all of the windage and friction torque, while the
DC supply provides all the electric and magnetic power loss. At still
higher speeds the average electromagnetic torque becomes negative
(generating). The average DC supply current remains positive until the
machine is going fast enough so that the generated electromagnetic
power exceeds the electric and magnetic power loss. This speed is the no-
load speed for generating. Between this speed and the original no-load
speed for motoring, the machine absorbs power from both the mechanical
and electrical "ports". At higher speeds, it becomes a pure generator.
The speed ranges are illustrated in Fig. 10.20 which shows the shaft
torque and the DC supply current. Motoring is when both are positive,
and generating is when both are negative. "Absorbing" is when the
torque is negative and the DC supply current is positive. The no-load

Fig. 10.20 Motoring, generating, and absorbing


10-34
10. E l e c t r o n ic C o m m u t a t io n o f S q u a r ew a v e M otors
speed for motoring is where the shaft torque is exactly zero, and the no-
loadspeed for generating is where the D C supply current is exactly zero.
The no-load speed for motoring is inherently lower than the no-load
speed for generating.

References
1, Miller TJE [1988] Switched reluctant* motor drives, PCIM Reference Book, Intertec
Communications, Ventura. California
2. Electro-Craft Handbook, Fifth Edition, August 1980, ISBN 0-960-1914-0-2
). Bolton HR and Mallinson NM [1986] Investigation into a class of brushle.ts DC motor
with quasisquan voltages and currents, IEE Proceedings, Vol. 133, Pt. B, No. 2, March
1986, pp. 103-111.

} Dote Y and Kinoshita S [1990] Brushless servomotors: fundamentals and applications,


Oxford University Press ISBN 0-19-859372-4
5 Kostenko M and Piotrowski L [1974] Electrical machines, MIR Publishers

10-35
11. PERFORMANCE EVALUATION BY TEST
11.1 Introduction
Many performance issues have been dealt with of necessity within some
of the previous chapters, in order to provide background and justification
for the rotor and stator design methods. However, several important
performance parameters must be considered more carefully in order to
facilitate a proper design for a particular application. For example, the
speed vs. torque curve has been mentioned, but a more detailed
presentation is given. This treatment leads up to the issue of sizing a
brushless motor at the outset before detailed design is performed. The
problem of heating is also related to the sizing, materials used, and
cooling methods. Temperature rise is a very difficult parameter to
calculate unless the thermal time constant has been measured on an
actual sample. Finally in this chapter, the so-called basic brushless motor
constants will be summarized to illustrate how the design is summarized
in terms of the performance specifications, why some of the motor
constants are desired, and how they characterize a particular motor.

11.2 Testing of PM brushless motors


The basic purpose of the speed vs. torque and current vs. torque plots is
to characterize the torque output capability of a particular brushless DC
motor at all possible speeds with a constant voltage applied to the phases.
Once this performance profile of a motor is known, it is possible to
design a power inverter/controller which will cause the motor to operate
within that speed vs. torque profile. Either voltage, current or both can
be used along with controlling the firing angles or commutation
positions of the rotor to achieve the desired performance for a variety of
applications.
It is a serious and time-consuming mistake to attempt to integrate the
design of a motor and power inverter/controller simultaneously without
first characterizing the motor by itself using a cons tan t-voltage 6-step
drive. This is also true of a brushless motor designed to be ultimately
driven with a sinewave drive. In fact, it will take less time to develop a
system if the motor is first tested with a constant-voltage 6-step drive
using very accurate commutation, without current limit.

11-1
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

The following parameters should be measured after the first samples are
built. A schematic diagram for each test is provided.
1. Back-EMF for each phase, using a synchronous motor to
drive the test motor and view the generated voltage on an
oscilloscope.
2. Resistance and inductance/phase.
3. Speed vs. torque & current vs. torque curve using x-y
plotter and dynamometer.
4. Thermal resistance.
5. Torque linearity with increasing load, to determine effects
of stator currents (armature reacdon).
6. Torque ripple at very low speed using a large inertia driven
by a 1-5 rev/min gearmotor.

11,2.1 Back-EMF testing


The set-up shown in Fig. 11.1 is perhaps the simplest and most useful test
which can be performed on a brushless DC motor. In fact, it is so useful
that it is perhaps the only performance test required on the production
line for motor qualification, because the relationship between the back-
EMF and torque is so predictable. If a simple commutation circuit is used
(Fig. 2.23), taking the sensor signals from the motor, the alignment
information is also available from this test as the back-EMF ripple will be
displayed on the oscilloscope. The flywheel inertia is not absolutely
necessary, but the results are better if it is used. An analysis of stator field
reaction is also available from the set-up shown in Fig. 11.1 with the
addition of a load and a current meter on the brushless motor output.
By adding a resistance as a load, current will flow, power will be
generated, and the voltage drop will indicate the reaction of the rotor
flux to the ampere-tums in the stator.
Further discussion of back-EMF testing will be found in Chapter 7.

11-2
11. P er fo r m a n c e e v a l u a t io n by t e s t

Fig. n .1 Back-EMF testing

Fig. 11.2 Measurement of phase resistance and inductance

11-3
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

1J.2.2 Resistance and inductance


The phase resistance can be measured using a precision ohmmeter, RLC
bridge, or multimeter, Fig. 11.2. Naturally the temperature of the
winding should be recorded at the time the resistance is measured. Any
discrepancy (more than 2 or 3%) between phases or between line-line
values may be an indication of an incorrect winding or connection.
The RLC bridge is also an easy way to measure the line-line inductance:
however, RLC bridge instruments generally use a test frequency of the
order of 1 kHz and very low currents, and the indicated inductance may
differ appreciably from the correct value which corresponds to a lower
frequency and much higher currents. The reasons for the difference are
the variable permeability of the core and the effect of induced currents
in the core and other parts of the motor. A complete experimental
analysis of the inductances requires DC inductance measurements using,
for example, the Prescot bridge described by Jones [1], which can be
modified to permit inductances to be measured with varying levels of bias
due to demagnetizing stator MMF or quadrature-axis stator MMF [2].
It is a good idea to take inductance measurements at several rotor
positions. In surface-magnet motors there should be very little inductance
variation as the rotor is rotated, unless there are problems with some of
the magnets.
In a three-phase motor, if the windings are connected in wye, the
measured line-line inductance is given by
L LL = Z, + - 21W l2 = 2 (L - M), (11.1)

where Lj = = L is the phase self-inductance and Af]2 = M is the


mutual inductance between phases. The measured line-line resistance is
R ll = 2R, i.e., twice the phase resistance, and the electrical time
constant (Chapter 14) can be taken as L ^ /R ^^ . This formulation can
also be used if the phases are connected in delta.
The line-line resistance and inductance are important if the motor is
operated with squarewave drive (two transistors on at any time). If it is
operated with sinewave drive, the parameters of interest are the per-phase
resistance R and the per-phase synchronous reactance both of which
are employed in constructing the phasor diagram, which is the main tool
11-4
11. P e r fo r m a n c e e v a l u a t io n by t e s t

for analysis of sinewave operation. Accordingly the appropriate methods


for measuring synchronous reactance should be used. The synchronous
j-eactance incorporates the self- and mutual inductances in a way that
correctly represents their effects with sinewave drive.
The synchronous reactance can be measured by a short-circuit generating
test, in which the motor is driven at low speed with all three phases
shorted together. The current is measured on at least one phase, and
preferably on all three. If the motor is wye-connected, the short-circuit
current /d(SQ is given by equation (6.90) with V= 0, and if resistance is
neglected,
^ ( 11.2 )
/ d(SQ

where E is the value of the open-circuit voltage per phase at the same
speed. The short-circuit current may well be much larger than rated
current, and may even be sufficient to partially demagnetize the magnets,
so this test should be conducted with caution. It is important not to apply
the short circuit suddenly while the machine is running, but to connect
the short circuit while the machine is stationary and then bring the speed
up slowly to the test value. This is because a sudden short-circuit
produces a higher transient current with DC offsets that may well
demagnetize the magnets.
With interior-magnet motors, i.e. salient-pole motors, it is necessary to
measure Aj and separately. The short-circuit test can still be used to
measure which replaces in equation (11.2), but X poses more of
a problem. The slip test normally conducted with wound-field synchro
nous machines may not be suitable with permanent magnet motors
because the excitation cannot be turned off and large currents would
flow when E and V were out of phase. These currents might
demagnetize the magnets. A simple method is to load the machine with
a three-phase resistive load following the short-circuit test, and measure
the current generated into the resistance, it is necessary to measure also
the phase angle y between and the current, and this can be done using
the motor encoder and an oscilloscope displaying the current waveform
along with the index pulse from the encoder, (The phase relationship
between E and the index pulse can be recorded in an open-circuit test).
Then equations (6.90-97) can be used to reconstruct the phasor diagram
anti extract a value for X .

11-5
D e sig n o f b r u s h l e s s pe r m a n e n t - m a g n e t m o t o r s

11.2.3 Speed/torque curve and load tests


The brushless DC motor would have very linear speed/torque and
current/torque curves unless reactance or inductance effects cause
otherwise. Therefore, the use of some simple or standard dynamometer
consisting of a load brake, torque/speed sensor outputs, current
transducer and an x-y plotter as shown in Fig. 11.3 can be used. The
brushless motor under test would be powered with a cons tan t-voltage
inverter with phase commutation control using angle position feedback
from the motor.
For torque measurement the most popular methods are to use an in-line
torque transducer or a load machine on gimballed bearings with a
torque reaction arm. Sometimes the motor under test is mounted on
gimballed bearings and the torque reaction arm is fitted directly to the
motor. Various types of load machines are used, including water brakes,
hysteresis brakes, friction brakes, magnetic particle brakes, eddy-current
brakes, and various combinations of AC or DC generators and load
banks. Many modem dynamometers are configurable so that the load
can be controlled via current, torque, power, etc., and automatic test
sequences can be programmed with all measurements (torque, speed,

Fig. 11.3 Measurement and plotting of speed and current versus torque
11-6
11. P e r fo r m a n c e e v a l u a t io n by t e s t

currents, voltages, power, efficiency, etc) taken automatically and


processed by computer virtually in real time.
Perhaps it is important to point out that for low voltage motors which use
ferrite or bonded NdFeB magnets, a series of speed/torque curves should
be analyzed. As previously mentioned, it is likely that the speed/torque
curve of such brushless motors will have a concave droop in the centre
of the curve between the no load speed point and the locked rotor
torque value. If the Hall sensor or encoder, whichever is used for
feedback commutation, is advanced or rotated opposite the direction of
rotation, an increase in torque and no load speed will result. If the motor
being designed and evaluated is to rotate only in one direction, a
significant improvement in performance can be achieved by utilizing
phase-advanced firing angles. This technique is particularly useful for 12V
fan and disc drive motors. A microprocessor can be used to control the
firing angles for start-up and at various speeds for either direction of
rotation in order to maximize performance. The facilitation of such
methods requires a careful characterization of the motor which can be
easily accomplished by plotting speed vs. torque curves for several phase
firing angles.

11.2.4 Thermal resistance


Fig. 11.4 indicates the required test set-up for measuring the thermal
resistance of a brushless DC motor. A calibrated thermistor must be
bonded to the winding end-turns using a thermal conducting epoxy with
the leads brought out for connection to a resistance bridge. The motor
is mounted to a dynamometer powered by a constant voltage drive. The
load is maintained at a value close to some expected or predicted
continuous rating until the temperature no longer increases. The
following data is recorded: input power to the motor, motor speed,
output torque, stator and ambient temperatures. The thermal resistance
is then calculated by subtracting the ambient temperature from the
steady stator temperature and dividing by the power loss in the motor
which is the input power in Watts minus the output power. The resulting
thermal resistance in C/W att can then be used in determining the
ultimate temperature rise of that particular motor for steady-state
operation if the losses are known or can be calculated.

It is standard practice to mount the motor under test to a plate of typical


thickness, size and material to act as a heat sink. There are some de facto
11-7
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

Fig. 11.4 Measurement of thermal resistance

standards for the plates in order to relate one motor to another. It would
be advisable to review the test conditions of the published data provided
by the competitor and use the same size plate for a heat sink.
The thermal resistance test can be performed with natural or forced
cooling, provided that the conditions are carefully recorded. Further
more, if the temperature rise of the winding is plotted as a function of
time, the initial rate of rise gives the thermal capacity of the motor.
Together with the thermal resistance, this can be used to estimate the
temperature rise under intermittent loading, as explained in Chapter 15.

11.2.5 Torque linearity


Sometimes the discovery of a nonlinear torque/current relationship (Aj.)
is a surprise because of the perception that the torque/Amp of a
permanent-magnet brushless motor is constant- There are three principal
causes of torque non-linearity in a brushless PM motor:
11-8
11. P e r fo r m a n c e e v a l u a t io n by t e s t
1. magnet heating;
2. armature reaction and saturation; and
3. inductance.
When a brushless motor is excited and accelerated rapidly to no-load
speed, very little current is drawn because the back-EMF nearly equals the
supply voltage. The motor remains cool and kp retains its maximum
value. As the motor is loaded, current is drawn and the motor heats up.
The remanent flux-density of the magnet decreases, and kj. decreases in
proportion.
Armature reactioni.e., the magnetic effect of stator currentdoes not
direcdy affect unless the magnetic circuit is saturated. If the magnetic
circuit is unsaturated, the magnetic flux set up by stator current is
superimposed on the magnet flux and kj, does not change. At high
current, however, saturation has the effect of decreasing Inductance
is another manifestation of the magnetic effect of stator current. Its effect
is more likely to be noticed at high speed. It slows down the commuta
tion of current from one phase to the next. By delaying the build-up of
current in an "incoming" phase, it decreases the torque during the first
part of the 60 conduction interval. Similarly, by delaying the extinction
of current in an "outgoing" phase, it can cause that phase to interact with
the next rotor pole in sequence and produce a "tail" of negative torque.
The reason why this effect is observed at high speed is that the
commutation intervals are fixed with respect to rotor position, but the
supply voltage and inductance are fixed, fixing di/dt. A finite rise-time
therefore spans a wider range of rotor movement as the speed increases.
Fig. 11.5 shows a typical test set-up for determining the torque linearity.
The same load dynamometer from the earlier tests can be used. The use
of an in-line torque transducer and a large synchronous motor or velocity
servomotor would be best, but a brake can be used if the loading can be
controlled to maintain a constant speed. A microprocessor is shown with
tachometer feedback. The voltage to the control must be increased from
some minimum to some maximum or set at specific values so that torque
and current measurements are taken with increasing current values. The
torque/Amp is then calculated for increasing currents while controlling
the load to maintain constant speed. An x-y plotter can be used to plot
torque vs. current and the curvature of the plot displays linearity. (See
also section 5.7.2 and Fig. 7.2).

11-9
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

Big. 11.5 Measurement of torque linearity

1 -5 RPM

Fig. ll.fi Measurement of torque ripple


11.2.6 Torque ripple
Torque ripple can be measured using a set-up similar to that shown in
Fig. 11.6. A very low speed gearmotor (1-5 rev/min) is used with a large
coupled flywheel driving through an in-line torque transducer. The
brushless motor is powered with a constant voltage inverter using a
11-10
11. P e r fo r m a n c e e v a l u a t io n by t e s t

sufficiently reduced input voltage so as not to overheat the motor and


drive. The torque transducer output is displayed on an oscilloscope. The
torque ripple can be analyzed including commutation errors caused by
inaccurate placing of the Hall devices or other sensors.

11.3 Magnetization testing


The testing of magnets is a specialized matter {3] but there are several
useful tests that can be performed by the motor manufacturer. One of
the simplest tests that is frequently necessary during assembly is polarity
testing. The safest way to do this is to observe the force of repulsion
between two similar magnets, suitably constrained, preferably by Plexiglass
guideways, so that the magnets cannot fly apart and cause injury, or fly together
and causefracture or chipping. Only the simplest of apparatus is needed for
this test. An alternative method is to offer the magnet up to a compass,

Fig. 11.7 Meaiurement of the degree of magnetization of a magnet


11-11
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

but this method should never be used because the compass needle is
liable to be remagnetized by motor magnets in an unpredictable
direction.
A very good and simple test for the degree of magnetization of a magnet
is shown in Fig. 11.7. The magnetized rotor is fitted into a close-fitting
steel cylinder which has a search coil wound in two semi-closed slots very
close together. As the rotor is rotated within the ring, the EMF generated
in the search coil can be integrated to give the flux passing through the
search coil, and this is a good measure of the remanent flux-density if the
rotor is a close push-fit within the steel ring. A Hall element can be used
instead of a search coil wound in slots. Alternatively, a Grassot fluxmeter
can be used with the search coil.

11.4 Precision dynamometer


Modern test facilities for adjustablc-speed motor drives can take
advantage of advanced data-acquisition technology and instrumentation.
Fig. 11.7 shows the configuration of a typical precision dynamometer [4], set
up for motor testing. Fig. 11.8 shows the same system extended to
include measurements on the complete drive system, and Fig. 11.9 shows
the general layout of this system.
Traditional dynamometers and associated instrumentation were set up to
measure sych quantities as mean shaft torque and speed, and RMS
currents and voltages, from which losses and efficiency and other
parameters of interest could be calculated. In modem adjustable-speed
drives the waveforms of these quantities are important as well as the peak,
mean and RAIS values, and it is often necessary to test not only over a
wide range of torque and speed, but also over a range of transient and
fault conditions. Inevitably this produces much larger quantities of data
than in the past, and computers are therefore an essential part of any
modern dynamometer system.
The transducers for measuring currents and voltages must not only be
accurate, but also they must have sufficient bandwidth to follow the detail
in the current and voltage waveforms especially under PWM operation.
Accordingly, flux-nulling or Hall-effect current and voltage sensors are
often used, and for very high bandwidth concentric current shunts are
preferred.
11-12
11. P e r fo r m a n c e e v a l u a t io n by t e s t

DRIVE SUPPLY
DATA
ACQUISITION
+
PROCESSING

T/S
POWER
LOAD ANALYSER
CONTROL TORQUE
SPEED T/S

XL
LOAD TEST
MACHINE T/S MACHINE ENCODER

Eg, 11.8 Precision dynamometer configured for motor testing

In Fig. 11.9 the data acquisition system has the capability of sampling
voltage or current waveforms at up to 10 million samples per second with
12-bit resolution (1 part in 4096). Waveforms with up to 64Ksamples can
be recorded, and many systems can record far larger numbers of samples.
It is important in power electronics and motor drive testing to have
differential inputs on measuring instruments, to avoid the problem of
single-ended input terminals (on oscilloscopes particularly, these are
usually grounded for safety).
Modem "wattmeters" are electronic, with the capability of measuring
voltage and current waveforms with harmonic content up to 400kHz or
more. They can typically provide peak, mean, and RMS readings of
voltage and current as well as mean and peak power, and some
instruments can even provide a harmonic analysis like a spectrum
analyzer.
11-13
D e s ic n o p b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

LOAD TEST
MACHINE T/S MACHINE ENCODER

fig. 11.9 Precision dynamometer configured for complete drive system testing
Electronic wattmeters can also be set up for the 2-wattmeter method, or
the 3-wire, 3-wattmeter method, or the 4-wire, 3-wattmcter method in 3-
phase systems. Such wattmeters are necessary especially when measuring
the separate efficiencies of the motor and its controller, because the
motor voltage and current waveforms generally are rich in harmonics.
Even with AC line current waveforms at the input to rectifiers etc., the
electronic wattmeter is a necessary instrument for power measurements.
The oscilloscope (Fig. 11.9) may not be an essential part of the
instrumentation, but is essential during setup to make sure that
everything is working properly and to help solve problems. Oscilloscopes
used in power electronics and motor drives work should have at least one
set of differential inputs, and digital storage oscilloscopes are preferred,
with single-shot (non-repetitivc) sampling rates of at least lOOMs/s. Many
digital storage oscilloscopes have optional waveform processing
calculators which can be used to determine peak, mean, and RMS values
and even to multiply waveforms together and calculate mean power.
11-14
11. P e r fo r m a n c e e v a l u a t io n by t e s t

Load re s is to r (fo r g e n e ra tin g te s ts )


C o n tro l c a b in e t/
m o to r c o n tro lle r

E le c tro n ic ^
w a ttm e te r s

Digital scope

Load c o n tro lle r

Load m a c h in e !
D a ta a c q u is itio n s y s te m

Tig. 11.10 Layout of precision dynamometer. Note the current probe amplifiers next
to the wattmeters; the inline torque transducer; and the emergency stop
button. The safety glass window at right permits tests to be observed from
outwith the dynamometer room when necessary.
The instruments and computer must be electronically linked, preferably
via the IEEE488 bus and protocol, or equivalent system. Software is
available from companies such as National Instruments and many of the
instrument makers (Hewlett-Packard, Tektronix, etc) for managing and
coordinating the instruments and for controlling the data acquisition and
processing. Any laboratory wishing to set up a precision dynamometer
should consult the suppliers of test equipment, measurement transducers,
and even complete dynamometers, for expert advice. Many engineers will
be amazed at the capability of modern instrumentation. Money invested
in good test equipment is well spent because it lasts for a long time and
gives quality information on the operation of prototypes and products.
To paraphrase Lord Kelvin, if you cant measure it you dont understand it.
11-15
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

References
1. Jones CV [1968] Unified theory of electrical machines, Butterworthl
2. Miller TJE [1981] Methods for testing permanent magnet AC motors, IAS Annual
meeting, Toronto, October 1981
3. Parker RJ [1990] Advances in permanent magrulitm, John Wiley and Som, ISBN ft-
471-82293-0
4. Staton DA and Miller TJE [1992] Validation of PC.-CAD Using a Precition DynamomtUr,
International Conference on Electrical Machines (ICEM), Manchester, 15-17 Sep
92,1221-1225

11-16
12. SIZING & COMPUTER-AIDED DESIGN
12.1 The modem design environment
The most usual situations which require a new brushless motor design
include a set of performance specifications and a physical size or
envelope. It is seldom that the performance requirement is given and
the size is open. In many cases a new performance specification must be
met through minor modifications to existing laminations and other
components, in order to avoid additional tooling costs.
When new designs are evolved from old ones, computer-aided design is
valuable in two particular ways:
1. Calculating and evaluating a large number of options,
often characterized by small changes in a large number of
parameters; and
2. Performing very detailed electromagnetic and mechanical
analysis to permit the design to be "stretched" to its limit
with confidence, while avoiding the need for a large
prototyping and test program, which would be expensive
and time-consuming.
Modern computer methods are rapidly reaching the stage where a new
prototype can be designed with such confidence that it will be "right first
time", without the need for reiteration of design and test that would
otherwise be necessary.
The computer-aided approach to design goes hand-in-hand with the
moder design engineering environment. Custom designs are increasingly
required within a very short space of time, while cost pressures force the
designer ever closer to the limits of materials and design capabilities.
Moreover, customers are becoming ever more sophisticated in their
requirements, and may specify (or ask to see) particular parameters that
traditionally were part of the "black art" of the motor builder. Often
these parameters are required for system simulation purposes long before
the motor is actually manufactured. Regulatory pressures on matters such
as energy efficiency, acoustic noise, and EMC also continually tighten the
constraints on the motor designer.
12-1
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

12.2 Basic sizing guidelines


No matter how effective the computer software available, it is always
important to check the overall parameters of a motor design using the
classical output equation which applies to (and unifies) all electrical
machines from the tiniest micromotors (a few pW) to the largest AC
motors used in process plants or ship propulsion (up to 20MW). This
equation is
T = KD?Lak (12.1)
where K is the output coefficient and D{ and /-stk are the rotor diameter
and stack length respectively. Two other coefficients sometimes used for
the same purpose are the torque per unit rotor volume TRV and the airgap
shear stess a. They arc both proportional to K, and the relationships
between them are as follows:

TOV = 4 r s<k ( 1 2 -2 )

Therefore TRV = 4K /n . The airgap shear stress o is the tangential


(torque-producing) force per unit of swept rotor surface area:
TRV = la. (12.3)

Hence
K = Z a = Z try (12.4)

n
The airgap shear stress o is measured in lbf/in or "psi". If Z>r and 1^.
are in inches, then T is in lbf-in. Note that an airgap shear stress of a =
1 lbf/in 2 corresponds to TRV = 13.8 kNm/m3. Typical values of K are
given in Table 12.1.
Electric and magnetic loadingsThe electric loading A is defined as the linear
current density around the airgap circumference:
^ _ Total ampere- conductors _ 2 /n A ^J ^ (12 5)
Airgap circumference trDr

12-2
12. S iztnc & C o m p u t e r -a id e d d e s ic n

where I is the RMS phase current, m is the number of phases, and Nph
is the number of turns in series per phase.
The magnetic loading B is considered as the average flux-density over the
rotor surface. In AC motors the flux-density is distributed sinusoidally so
that the fundamental flux per pole is given by
$ - B * wD' L* ( 12 .6)
1 2p

3tid then the generated EMF per phase is


B = -1 V. (12.7)

The maximum available airgap power is mEI - Ta>/p, assuming 100%


efficiency and the correct orientation of the flux and MMF, and from
these equations the torque per unit rotor volume can be determined as
= KB A N m /m 3 (12 -8 )
K
wilh
K = BA = 1 .7 4 k-.,BA 1.6BA (1 2 .9 )
v/2 *
if is approximately 0.9. Correspondingly, TRV ~2 BA and o = BA. For
example, with an electric loading of 20 A/mm (i.e., 500 A /in) and a
magnetic loading of 0.5 T, o = 0.5 x 20 x 10s = 10 kN /m 2 = 1.45 lbf/in2.
The interaction between flux and current in the production of torque is
implicit in the formula o = BA. Note that B and A are both densities,
respectively of flux and current It is useful to understand how these
densities affect the size and shape of the machine. In conventional
motors the flux crosses the airgap radially and it is natural to evaluate the
magnetic loading as the average radial flux-density in the airgap. Its value
is limited by the available MMF of the excitation source. The MMF
required depends on the length of the airgap and the saturation
characteristics of the rotor and stator steel. In most motors, if the flux
density in the stator teeth exceeds about 1.8T then the excitation MMF

12-3
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

requirement becomes too great to be provided economically, because of


the sharp increase in H in the steel. In slotted structures, the flux
density in the teeth is typically twice that in the airgap. Therefore, the
flux-density is normally limited to a peak value around 0.9T, which gives
B - 0.57T. Saturation of the rotor or stator yoke (back-iron') is a further
potential cause of excessive excitation MMF requirement.
The electric loading is a question of how many amperes can be packed
together in each unit of stator circumference. Basically this is limited by
the slot fill factor, the depth of slot, the current density, and the cooling.

Typical values of 0 , K, TRV


o lbf/in2 K lbf-in/in3 T R V k N in /m
Small totally-enclosed motors 0.5 - 1 0.8-1.6 7 - 14
(Ferrite)
Totally-enclosed motors 1-3 1.6 - 4.7 1 4 -4 2
(Sintered Rare Earth or NdFeB)
Totally-enclosed motors 1.5 typ. 2.4 typ. 21 typ.
(Bonded NdFeB)
Integral-hp industrial motors 0.5 - 2 0.6 - 3 7 - 30
High-performance servomotors 1- 3 1.5 - 5 15 - 50
Aerospace machines fc . 3 - 7.5 30 - 75
Large liquid-cooled machines 1 0 -1 5 15 - 200 100 - 250

Guide values of o (lbf/in2)


Low <1
Medium 1- 2
High >2
T able 12.1 G uide values fo r airgap shear stress , trv , e t c .
It is interesting to see why it is the rotor volume and not its surface are?
that primarily determines the torque capability or specific output. As
the diameter is increased, both the current and the flux increase if the
electric and magnetic loadings are kept the same. Hence the diameter
12-4
12. S iz in g Sc C o m p u t e r -a id e d d e sig n

(or radius) appears squared in any expression for specific output On the
other hand, if the length is increased, only the flux increases, not the
current. Therefore the length appears linearly in the specific output.
Thus the specific output is proportional to D ^ L ^, or rotor volume. In
practice as the diameter is increased, the electric loading can be
increased also, because the cooling can be made more efficient without
reducing the efficiency. Consequently the specific output ( t r v ) increases
faster than the rotor volume.
For totally-enclosed motors the lower values of o, TRV, and ATwould apply
with natural convection, while the higher values would apply with forced-
air cooling supplied by an external or shaft-mounted fan.
It is of interest to relate the electric loading to the current density in the
slots. With a slot depth of 15mm, a gross slot fill factor of 40%, and a
tooth width/pitch ratio of 0.5, the current density is
J = ------------------------------- ---------------------------------
Slot-fill x Slot-width/slot-pitch * Slot depth ( 12. 10)
= --------
0.4 x 0.5--------
* 15 = 6.7 A/nun2.
Typical values of current density for use in different applications are
given in Table 12.2.
Condition A/mm A /in2
Totally enclosed 1.5-5 1000-3000
Air-over 5-10 3000-6000
Fan-coolcd
Liquid cooled 10-30 6000-20000

TABLE 12.2 TYPICAL CURRENT DENSITIES


These current-density values assume that the stator windings are heavily
varnished for good heat transfer. In the case of the air-cooled densities,
ihc fan is mounted on the rear of the motor outside the frame with a
shroud which focuses the air over the OD of the motor with the higher
values relating to "air over" (AO) designs with fins on the housing. The
12-5
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

liquid cooled motors would have a passageway around the entire outside
diameter of the stator with a cooling fluid circulating to remove the heat
The higher numbers would be appropriate to motors in which, in
addition to the outside cooling jacket, there would be tubes up and down
the slots potted in epoxy. If hollow conductor wire is used with cooling
fluid circulating inside, no potting is necessary. (This is called "direct
conductor cooling").
When a variety of motor types and sizes of machines are evaluated to
determine their airgap shear stress at their name plate ratings, a broad
range of values is encountered, with variations as much as 2000% for a
given motor type. It is, therefore, difficult to use this method of "sizing"
a design unless a particular airgap shear stress figure can be associated
with a particular method of construction and cooling. The permanent
magnet brushless motor will exhibit a more consistently predictable
airgap shear stress than most other types of motors, because for each
different magnet grade the magnetic loading is fixed within fairly narrow
limits, while the method of construction and cooling is reasonably
consistent among different motors.
After the magnet grade has been selected and the first estimate made of
the rotor diameter and active magnetic length, the stator OD can be
estimated. For interior-rotor motors, the rotor diameter can be divided
by 0.45 to 0.50 to arrive at a stator lamination OD. For exterior-rotor
motors the rotor OD is determined by adding twice the magnet thickness
plus twice the rotor cup thickness to the rotor diameter Dr. These
dimensions are calculated according to the practice described in earlier
chapters.

12.3 Computer-aided design with PC-BDC


Most of the design equations developed in this book have been
programmed into the PC-BDC computer program at the University of
Glasgow SPEED Laboratory, and a brief description of the main facilities
of this program will illustrate the scope of a certain type of GAD program
for motor design.
Fig. 12.1 shows the opening MAIN MENU of PC-BDC, listing (on the left)
the main tasks for which the program can be used. Each of the options
or tasks covers a wide range of design engineering tasks.
12-6
12. S iz in g & C o m p u t e r -a id e d d e s ig n

rpcHDC 4.UU: Mail! menu (C :\PCBDC4 .UNfllKCfil1.BD4) 86K f5 1


PC-BBC Main T itle 23rd Hay 1994
PC-BDC Sub-title 16:2?
1 Hotor crDes-Bectlon editor
2 Static daslgn
p flynawlc design___________________
3 [ -
[iTnp^T.lo. -jJilor
4 Output design s)int
~
5 Steal database
6 Kignat dttalmee
H Hanranle ana Lye Is
T Speed/Tarque characteristic
C Display siwlatlan result graphs
U Winding design
t Print laninations and graphs
8 User preferences
F File translation
Q Return to DOS

Option 7...

(c) TJE Hiller,HI HcCllp.DA Staton. SPEED Laboratory. Glasgow University. 1994
Licensed to Prof. TJE Miller
||nlp: Choose an o p t i o n by p r e ssing thu h iyh I iyiited ken.

Fig. 12.1 Main menu of the PC-BDC computer program


Fig. 12.2 is a simplified structural diagram of the program, which shows
that the program is not really a design synthesis program, but is really a
specialized design calculator. The user defines the design within certain
constraints as to type of rotor, selection of materials, winding layout,
control strategy, etc., and then the program calculates the performance
under the conditions defined by the user. Refinement of the design is
done by the user, making changes to the motor dimensions, turns, wire
size, control parameters, etc., as he/she deems necessary based on an
assessment of the calculated performance figures. The computer program
provides no clues as to what changes should be made in order to move
closer towards a given design objective, and therefore all the creative
judgement lies with the user. In this sense PC-BDC is quite the opposite
of a knowledge-based design synthesis program, the concept of which is
to replace the designer and do his job. The concept of PC-BDC is not to
replace the designer but only to relieve him /her of repetitive calculations
and at the same time, to provide vastly more data for each design than
the designer would normally have time to calculate. (Over 500
parameters are calculated for each design, although normally only about
240 are displayed in the output) . The calculation capability of PC-BDC is
constantly being refined and extended to newer motor configurations.

12-7
DESICN o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

Fig. 12.2 Structure of PC-RDC. from the user's viewpoint


Starting a new designThe program provides several methods for starting
a new design. There is a set of "hot" keys which recalls a number of
standard motor designs, which can subsequently be edited to modify the
performance. Alternatively, a new design can be started by filling in a
blank template of dimensional and winding data. A third method is
through the simple sizing procedure described at the beginning of this
chapter. The entry panel for this method is shown in Fig. 12 3. Once the
basic requirements are entered in this panel, PC-BDC returns with a
suggested cross-section, as shown in Fig. 12.4.

12-8
12. S iz in g & C o m p u t e r -a id e d d e s ig n

PC-BDC 4,00: Hu t u r sizing procedure___________ B 6K f5

This procedure Enables you to got nan design started ulth


appraxlmtely the right ehipo And size. Once the design Is
etartBdj use Options 1 and 3 to Modify It.

[Torque............... . .Output of motor in Nn J 1.0BB8


TAr...... . .SpBdflc torquB / rotor ijolune In VN*/*3 10.0600
L/D ratio. , ..Ratio of rotor length *nd dlaneter 1.0608
Poles..... . .Nunber of poles 4
Slots..... . .Nunber of slots 24
Us........ ..Supply voltage 24.8BB8
m ....... ..Operating speed zeee.oEee
Rotor type. SurfRad

Help; F2 Edit FAB Kent field Shift<Tgb frev field Ctil*S Save ESC Exit

Fig. 12.3 Entry panel for sizing procedure in PC-BDC

The cross-section in Pig. 12.4 can be edited and re-drawn according to


requirements. The ability to make changes and view them graphically is
extremely important in developing a visual association between particular
cross-sectional features and corresponding performance characteristics.
For example, it is generally immediately obvious whether the design is
intended for rare-earth magnets or ferrite magnets, or whether it is a
sinewave or squarewave motor. The distinguishing features might be
obvious to experienced designers, but the program is invaluable as a
training tool because of the speed at which less experienced engineers
can become acquainted with them.

Fig.12.5 shows the winding editor which can be used to modify or build
any form of winding. The program can automatically build concentric,
lap, and fractional-slot windings, but it also provides a facility for
insertingcoils entirely at the users discretion. Regardless of the number
or disposition of the coils, the program can calculate the back-EMF
waveform, the resistance, self-and mutual inductances, and the
performance with squarewave or sinewave drive.

12-9
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

fig. 12.4 Gross-jcction editor in PC-BDC


It is even possible to simulate unbalanced three-phase windings, which
are occasionally used. With windings connected in delta, the program
displays the loop EMF in the delta along with the main phase and line-
line back-EMFs, so that the design can be checked for circulating third-
harmonic in the delta.
PC-BDC can calculates the performance with ideal squarewave or sinewave
currents, and it can also perform a time-stepping simulation with any one
of six different control strategies in squarewave mode. Several options are
provided for tailoring the program to user requirements, including the
ability to customize the output design sheet and to calculate user-defined
parameters.
PC-BDC incorporates databases for permanent magnet materials and
electrical steels, which are easily edited or extended by the user.
Fig. 12.6 shows the template editor for editing parameters not in the cross-
section, including winding parameters and controller parameters. Figs.
12.7 and 12.8 show typical waveforms, and Fig. 12.9 shows typical output
data.
12-10
12. S iz in g & COMPUTER-AIDED DESIGN

Fig 12.5 Winding editor in PC-BDC.


I FC-MM/ 4.UU:TufipUte Editor CCAPCBDC1!.0\CH12.BD4) 1H2K f5 1
OiMniwional Parameters
iifm 9.061 Radi 25.154 Gap 0.503 LH G .907
Ba'.iM 150.000 POLES 4 Had3 48.008 SLOTS 24
RHSQ Squire Tm 2.768 SltDpth 12.074 S ItOpen 0.92B
SltODpth 0.700 SltOflng 49.000 Lstk 5B.3B8 RotType SurfHad
Inset 1.000 Width 2.000 M1H 0.080 Stf B.97B

Uinding Parameters
DdgTypa Lap Throw 5 Co i1E/P 1 TC 10
HSH 1 PPftTHS 1 Ext B.000 Liner B.4B8
WireSpec BareDIo Wire 1.226 Skew B.60B

Control Parameters
KPH 2006.006 Conns* 3-Ph Uua Idaus Square Su Ctl C120 Q6
ISP 10,990 DuCy 1.000 CFrq 5.BBS Us 24.B00
0* 1.000 If b .b b s IM 0.600 THB 9.000

fltfcer Parameters
IUU n ita tm UdgTemp HagTenp
bDDDnm
25.000 25.0B8 Ik B.BBB
Uf0 0.006 RPMB oqqq UUU HUFT 2.000 XLph 1.060
XFs 1.000 Xrl 1.900 XET 1.008 XBtpk 1.0BB
TenpCalc DegCU DegCU 0.100 HTranflct IB.BBS HTranEnd 6.0GB
flubSent 20.000 EndFI11 0.500 1Of \
| Helji: Shaft radius Fl-Hulp |
Fig. 12.6 Template editor for input data in PC-BDC
12-11
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a c n e t m o t o r s

yw x 1.0-0
<^WMyyVWwvH|yvvwwvv\yyvvvvvm^y^

Fig. 12.7 Typical EMF, current, and torque waveforms in PC-BDC


PC-BDC 1,00: Des i g n Sh e e t r^K f5
4 b l i n di n g D a t a s---------------------------
UdgTupe Lip Connex Ph Uye Throw 5
TC 10 ColIb /P 1.060 NSH 1
Tph 4a.eaa PPflTHS 1 SPP 2.000
Layers 1.000 CSidesPh B Z 24B.0S0
KLT 240.107 m LgthOEnd 94.735 m Ext 0.000 m
EndF i11 9.500 LaxPack 57.566 im TTRho 1.020
SFlIl 0.190 Ulra 1.226 UlreSpec BareDU
UlraDla 1.22E m Liner 0.400 m
SlotRrea B2.127 m *2 Condflrea 1.180 Mra2 RSIotU 49.165 wT2
UdgTenp 25.600 DegC Rph 0.141 Dhirth Stern 6.282 Dims
Lph 0.272 pfl/Ph Hph -0.B57 ftVP h Ltern 0.430 bH
0.142 rfi/Ph LS lot B.B99 wH/Ph Lsndt 0.931 nH/Ph
Lb
Hg -0.057 nH/Ph HSIot 0.600 rH/?h fz 1.149
PCS lot 1.955 XLph 1.600 XET 1.860
Ak I 52.500 nSag

5 Magnetic Circuit Design:

BrT 0.401 T BgOC 0.289 T HcT 297.950 kA/n


BglDC 0.355 T PhlHl 0.458 Mb PhlC 0.487 NUb
H E M * 8.378 T Bn/BrT 9.942 f Lkg 0.850
| He 1p : Average uien-circuit f lux-density in the magnet FI Help
r.g- 12.8 Typical output data from PC-BDC
12-12
12. S iz in g & C o m p u t e r -a id e d d e s ig n

| K-BDC 4..CIO: Design Sheet 74K f5 |


7 PH Dynamic design (tine-stepping simalatlnn);---- - --------- ------------------
OpHode Motoring
Torqua a.585 Nn PcwerSh 122.5B3 U Eff 80.166 x
loosest 20.150 U loasFa_S 2.151 U LosaUF_S 0.06B M
LossTot 38.389 U Tempiiso 3.031 DegC Jrn 6.911 A/W2
IUpk 11.107 A lUav 6.686 A [Urns B .154 A
ILpk 11.107 ft llav 6.686 A ILrns 8.154 A
[Qclipk 11.107 ft IQchav 2.421 A IQchrns 4.917 A
IQcnpk 11.107 A IQcnav 3.311 A IQcnrra 5.774 A
IDchpJf 11.107 A IDchau 6.B92 A IDchrna 2.BB1 A
[Dcnpk 8.800 A IDcnav B.0B8 A IDc h t r s 0.080 A
[DCI iiikP 7.151 A Lo s b Co d u 18.003 U EffDCSh 71.382 x

B Miscellaneous:---------------

IftCu 0.299 kg UtFa 1.E90 kg UtHag 0.191 kg


tilTot 2.1B1 kg RotJ 1.5471-04 kg-2 LosFe/Ut 1.272 U/kg
[JCLinkU 8.136 A sigM B.424 psl KFb 1.308
Frftql 6G.667 Hz FroqChop 4.BBS kHz
lc DsgCU DagCU 8.188 dsgC/U HTranAct 659.996 U/M2/C
flwhloot 28.808 DagC HTrnnEnd 0.B88 U/2/C
HijsSjud 12.588 x IntStep 0.5H0 eDeg
Help; (JidtJi of ligsteresis band as a percentage of ISP FI Help |

Fig. 12.9 Typical performance data from PC-BDC

12.4 finite Element Analysis


12.4.1 Introduction
The finite-element method is a numerical method for solving
electromagnetic field problems which are too complex to be solved using
analytical techniques, especially those involving non-linear material
characteristics. The method basically involves the discretization of the
motor cross-section (or volume in the case of 3-D analysis) into smaller
areas/volumes called finite elements. The spatial variation of magnetic
potential (vector or scalar) throughout the motor is described by a non
linear partial differential equation derived from Maxwells equations. In
its linear form this equation is Laplaces equation, Poissons equation, or
the Helmholtz equation, depending on the type of problem being solved.
Usually the partial differential equation is written in terms of the vector
potential A, because of the economy with which important field
quantities such as flux-density, flux, etc. can be subsequently determined.
Accordingly the partial differential equation is solved, after discretization,
12-13
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

in terms of A, and the other "output" quanddes are calculated from the
nodal values of A in the "postprocessing" phase.
Within one element, the vector potential is assumed to vary according to
a simple shape function, which may be a linear, quadratic, or higher-
order function of the three sets of node coordinates for the vertices of
the triangular element. Linear elements give the fastest solution but the
least smoothness in the field variation.
Modem finite-element software is extremely robust and accurate, and has
innumerable features to assist the user to set the problem up and extract
the engineering parameters of interest. However, the point should be
made that the finite-element method differs from rapid-design software
like PC-BDC in two important ways:
1 . The finite-element method is generally limited to one specific
type of problem, such as electromagnetic or thermal analysis,
whereas rapid- design software calculates a wide range of
parameters ranging from weights and inertias through to
performance and dynamic waveform calculations, temperature
rise, and many non-electromagnetic parameters.
2. The finite-element method is intended for accurate analysis,
and the emphasis on accuracy means that setup and execution
time is much longer than with rapid-design software. A typical
simple finite-element exercise might take days or weeks, but a
complete motor can be designed in less than one day with a
program like PC-BDC. Inevitably, PC-BDCs electromagnetic
calculations cannot be expected to be as accurate as those of the
finite-element method, although the difference in accuracy is
often too small to be of concern. On the other hand, many
detailed problems such as the calculation of cogging torque,
which are beyond the scope of rapid-design software, may require
extremely lengthy finite-element calculations lasting several weeks.
Application of the finite element method to machine design involves
three stages:
1. Pre-Processing
2. Field Solution
3. Post-Processing
12-14
12. S iz in g & C o m p u t e r -a id e d d e s ig n

12.4.2 Pre-Processing
In most cases this is the most user-intensive part of finite-element
analysis. Three tasks must be performed:
1. Mesh Generation
2. Material Definition
3. Problem Definition
Mesh, generation involves division of the motor cross-section into a set of
triangular elements (2-D solutions) or division of the motor volume into
'bricks' (3-D solutions). Modem mesh generation is carried out using
either interactive graphical techniques (using conventional CAD drafting
software such as AUTOCAD, or using the internal specialist drafting
facilities of the finite-element software itself. Another alternative is the
use of using specialist mesh generation code written in a high-level
language or command file.
Interactive graphical mesh generation is the quickest way to form a finite-
element mesh in the majority of cases. However, a drawback of using this
technique is that when it is required to calculate a series of solutions for
motors of the same general type, but with different dimensions, it will
usually be necessary to generate the mesh individually from the
beginning for each case. This problem can be overcome by writing
specialist mesh-generation software, which may be in the command
language of the finite-element program, or alternatively in C, PASCAL or
FORTRAN. The interface to the finite-element program may m e a
standard data format such as .DXF or a format specific to the particular
vendor. This method of mesh generation is especially useful when
generating a range of meshes for motors of similar type but having either
different dimensions or different rotor positions.
Specialist mesh generation software calculates the coordinates required
to define the motor geometry. The cross-section is usually split up into
regions representing different "materials" such as current-carrying
conductors, air, steel, and magnets. Each region may define a different
component used in the construction of the motor, for example, the shaft,
rotor core, magnets, stator lamination, airgap, etc. In most cases it is
beneficial to split the components further into smaller polygons along
lines of symmetry.

12-15
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a c n e t m o t o r s

For example, a stator lamination can be created by reflection followed by


multiple rotational copies of half a slot pitch. This procedure reduces the
amount of data needed to specify the geometry, and reduces the chance
of errors.
The coordinates of vertices defining region boundaries must be
calculated from dimensional parameters such as the outer frame
diameter, slot depth, tooth width, etc. Usually this involves a "translation"
calculation from the engineering parameters to the finite-element model.
Mesh grading constants must also be specified. These determine the
grading of the node-spacing along each edge of the polygonal regions.
When rotation of the motor rotor is to be modeled, it is essential to
define the airgap using a sliding surface, splitting the airgap into at least
two layers. One of these layers is fixed to the rotor and one to the stator.
12-16
12. S iz in g & C o m p u t e r -a id e d d e s ig n

Fig. 12.11 Finite-element mesh for one pole-pitch of a. brushless DC motor, showing
the use of periodic boundary conditions
The node spacing on the central sliding surface is set to a constant such
thatit is possible to rotate the rotor by any multiple of this constant. Fig.
12.10 shows a mesh in which the airgap region is divided into four layers
and the sliding surface is central to the airgap. Fig 12.11 shows the full
mesh for one pole-pitch of the same motor.

An alternative approach involves developing algorithms in which the


stator bore and rotor outer diameter nodal coordinates are calculatcd
and the airgap elements are so defined as to give the best possible shape
(as near as possible to equilateral triangles) [ 1 ].
Material definition involves curve fitting the non-linear B-H characteristics
of the steel and magnet materials used in the construction of the motor.
Many finite-element programs include databases of material data for the
convenience of the user.
12-17
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

Fig. 12.12 Leakage flux paths in n "spoke" type brushleu permanent-magnet motor
Problem definition involves the application of the correct boundary
conditions, imposing the correct current densities in the appropriate
winding elements, and definition of the direction of magnetization of
magnets. Periodic boundary conditions should be used if possible, as
they make it possible to model only a fraction of the cross-section (Fig.
12.11). For example, in the case of a 4-pole motor, if the number of
slots/pole is an integer, it is only necessary to model the motor over one
pole pitch.
12.4.3 Field Solution
The solution of the discretized partial differential equation uses
specialized mathematical algorithms developed over many years [4]. The
algorithm is often based on the minimisation of an energy functional,
that is, a mathematical function that is related to the stored potential
energy in the field.
12-18
12. S iz in g & C o m p u t e r -a id e d d e s ig n
The discretization transforms the partial differential equation into a large
number of simultaneous nonlinear algebraic equations containing the
unknown node potentials. Iteration is essential and the Newton-Raphson
and congjugate-gradient procedures are widely used. With linear
elements, the potential is assumed to vary linearly between nodes and the
flux density is constant within each element. Current density is also
assumed to be constant within each element associated with a winding.
12.4.4 Post-Processing
The field solution is in terms of magnetic vector or scalar potential, but
the design engineer needs quantities such as flux densities, force and
torque. The extraction of these quantities from the potential solution is
called post-processing. A good interactive graphics facility is important for
SO that the essential information and parameters can be extracted from
the large number of node potentials effectively and quickly. Finite-
element analysis can be used to generate the following output:
1. Flux plots. These are especially useful for forming a picture of the
flux. They can also be used for estimating leakage flux and
calculating leakage permeanccs. Fig 12.12 shows a flux plot for a
spoke type motor, in which the finite-element method is useful for
calculating rotor leakage flux.
2. Fhtx calculations. The flux between two points of interest is
calculated from the difference in vector potential at the two
points, multiplied by the axial length. This calculation can be
extended to obtain flux-fiwAagtf, and hence inductance of windings.
8. Fhix density contours. Coloured filled zones can be used to indicate
areas of high local saturation. Flux density values at any point can
be readily obtained by using a cursor.
4, Graphs offlu x density variation. The required component of flux
density (radial, x-component, y-component etc.) can be plotted
along a pre-dctermined path. For example, the variation in radial
component of airgap flux density can be plotted around the rotor,
and examples are shown in Figs. 12.8 and 12.19.
Flux density vector plot A field of arrows is plotted over the cross-
section, representing the local flux density vector.
12-19
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

6. Magnet working point distribution plot. The B /H working point


within each magnet element can be superimposed on the major
B /H curve of the magnet. This is useful for visualizing the effect
of demagnetization.
7. Back-EMF and con-lass calculations. Finite-element analysis can be
performed for incremental rotor positions from which variation
of flux and flux-linkage against rotor position can be obtained.
This can be used to calculate back-emf (by differentiation with
respect to time) and core losses.
8. Force and torque. Force and torque are commonly calculated using
the following three methods; Lorentz force, Maxwell stress and
virtual work. All three are prone to errors [3], and the best
method depends upon the type of problem and the users
preferences.
9. Inductance. Self and mutual inductances can be calculated from
the stored energy or flux-linkage [4], However, the calculation is
made more complex due to the presence of magnet flux. One
method sometimes used to overcome this problem is to carry out
a field solution analysis at a particular value of load current. The
permeability values from this solution are stored and used in a
second solution which has the magnets removed.
While there are many finite-element programs on the market, the
general form of the pre-processor, solver and post-processor does not
vary greatly from one program to another. In this discussion all results
were obtained using the package OPERA-2d [5]. To illustrate the use of
finite-element analysis in machine design it might be useful to include an
example. Section 12.5 describes the analysis of armature reaction in
brushless DC motors using finite-element analysis.

12-5 Example: armature reaction in brushless DC motor


Permanent-magnet motors produce torque by interaction between the
magnet flux and the current flowing through the armature windings. The
armature current sets up its own magnetic field which distorts the airgap
flux distribution. The effect is similar to that in DC and synchronous
machines and is known as 'armature reaction [6].
12-20
12. S iz in g & C o m p u t e r -a id e d d e s ig n

While the magnet flux remains substantially constant, the field set up by
the armature windings is proportional to the current. Motors are usually
designed such that the armature reaction does not affect the torque per
ampere, ky. Under heavy load conditions armature reaction is liable to
decrease kT.
Even under normal load conditions, the distortion in the airgap flux-
density distribution tends to increase the core losses. It is not unusual for
the core losses at rated load to be double the no-load value, even though
Jfcf may not vary significantly between no-load and full load. This is
because depends on the magnet flux alone, whereas the core losses
depend on the total airgap flux and its distribution.
The effects of armature reaction can be classified into 9-axis effects and
tfaxis effects:
1. f-axis : cross-magnetization or distortion
2.d-axis : demagnetization or reduction of airgap flux density.
These effects are illustrated with the help of flux plots and flux density
graphs generated using finite-element analysis. A small squarewave
brushless DC motor with four poles and twelve slots is used as an
example.
The examples can be understood by looking first at the flux distribution
set up by the magnets alone, then at the flux distribution set up by the
armature current alone, and finally at the superposition of both, keeping
in mind the fact that the superposition is non-linear due to magnetic
saturation effects.
12.5.1 Open-circuit flux distribution

Fig. 12.13 shows the flux distribution created by the rotor magnets
alone. This is also referred to as the open-circuit flux distribution since
the armature or stator windings are not excited in this case.
Many important features are immediately apparent in Fig. 12.15, for
example, the radial magnetization of the magnets, the concentration of
flux in the teeth, and the variation of flux-density in the stator and rotor
yokes.
12-21
D e s ic n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

Fig. 12.13 Flux distribution due to magnets acting alone (open-circuit condition)
Closer examination shows more subtle effects such as the fringing around
the slot openings, discussed at length in Chapter 8. The high coercivity
and low recoil permeability of the magnets is the reason for the extreme
regularity of the field lines in the magnets. "Hard" magnets such as
ferrite or rare-earth magnets act as "rigid" sources of flux, and there is
very little distortion of the field inside the magnets due to slotting.
12.5.2 Armature reaction field alone
Fig. 12.14 shows the flux distribution set up by the armature currents
acting alone, assuming two-phase-on operation with a severe overload.
The airgap as seen by the armature includes the magnet thickness and
hence appears as a high-reluctance path for the armature field. The
magnets are considered to be removed or unmagnetized, and have no
effect on this distribution. Note that the flux encircles the ampere
conductors in the manner of Amperes Law, and appears in four loops.
12-22
12. S iz in g & C o m p u t e r -a id e d d e s ig n

Kg. 12.14 Flux distribution due to 'armature current acting alone (two phases on).
The dotj and crosses show the polarities and locations of the armature
currents. There are approximately 1600 Ampere-conductors per slot, with
a high value of current density (17.3 A/mm2).

A proportion of the flux in each loop crosses the long airgap into the
rotor yoke twice. Most of the remainder crosses the tooth-tops, while a
small fraction travels circumferentially around the space left by the
magnets. These separate components of flux are estimated by simple
formulas in Chapters 5 and 6, in the calculation of the self-inductance of
ihe winding and the mutual inductance between windings. The
somewhat complex shape of the flux paths shows that this type of
calculation can only ever be an approximation, although surprisingly
accurate results are often achieved. With finite-element analysis, the
inductance can in principle be calculated more accurately.
Note that all the flux-plots shown here are two-dimensional, and three-
dimensional effects (end-effects) are neglected.
12-23
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

Fig. 12.15 Resultant flux-diatribution with magnets and cross-magne tizing armature
reaction. The currents are as defined in Fig. 12.14.
The inductance associated with the flux pattern in Fig. 12.14 is a static
inductance, which is directly useable in time-stepping simulation of
squarewave or sinewave brushless motors, but is quite distinct from the
synchronous reactance discussed in Chapter 6.
The synchronous reactance cannot be directly calculated from a static
finite-element flux-plot, except in the special case when the windings and
the flux are perfectly sine-distributed. In the general case, the
synchronous reactance calculation requires the stepping of the rotor
through at least one electrical revolution, followed by a differentiation of
the flux-linkages of the phase windings and a harmonic analysis to extract
the fundamental component. This is because the synchronous reactance
is defined for a fundamental space-harmonic field so that it can appear
correctly in the phasor diagram.

12-24
12. S iz in g & C o m p u t e r -a id e d d e s ig n

Values qIB M O D
___ V bIu m o IBM OD

Fig- 12.16 Variation of airgap flux-density under one pole. Dotted line : magnets
alone. Solid line : resultant with cross-magnetizing armature reaction.

12.5.3 Cross-magnetization
The cross-magnetization effect is greatest when the rotor magnet field is
in quadrature with the armature field, as in Fig. 12.15. Many significant
features of the effect of armature reaction can be observed from Fig.
12.15, most important of which is the increase in flux-density at the
leading tip of die magnet pole and the decrease in flux-density at the
trailing tip. This effect is illustrated in Fig. 12.16, which shows the
variation of airgap flux density under one magnet pole around the
airgap. The dotted line corresponds to Fig. 12.13 showing the variation
in flux density due to magnets alone. The solid line corresponds to Fig.
12.15 showing the resultant distorted flux distribution when both
magnets and armature current are present.

12.5.4 Demagnetization
The demagnetizing effect is greatest when the axis of symmetry of the
magnet (the <axis) is aligned with the axis of symmetry of the armature
12-25
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

Fig. 12.17 Flux distribution due to magnets acting alone. The rotor is turned 45*
clockwise relative to its position in Fig. 12.13
ampere-conductor distribution. Fig. 12.17 is again the open-circuit flux
distribution due to magnets alone, but the rotor is rotated 45 degrees (90
electrical degrees) to align the two axes as required.
The armature ampere-conductor distribution remains unchanged from
that of Fig. 12.14, and the resultant flux distribution is shown in Fig.
12.18. The airgap flux density at the centre of the magnet is much less
than the open-circuit value in Fig. 12.17. The MMFs of the magnet and
the armature current combine in the region of the quadrature axiSj
where the flux is forced to return from one magnet pole to the next
across the tooth-tops, rather than following its normal course through the
teeth and round the stator yoke. The stator yoke flux is gready reduced.
The armature reaction in all these examples is for very high current,
corresponding to locked-rotor condition.
12-26
12. S iz in g & C o m p u t e r -a id e d d e s ig n

Fig. 12.18 Resultant flux distribution with magnets and demagnetizing armature
reaction. The magnet (taxis is aligned with the axis of symmetry of the
armature ampere-conductor distribution, such that their MMFs are in
opposition

In Fig. 12.19 the dotted line shows the variation of flux density around
the airgap with the magnets acting alone. The solid line shows the flux
density variation when both magnets and armature current are present.
Note that both the cross-magnetizing and demagnetizing components of
armature-reaction MMF distort the airgap flux waveform and introduce
additional time-harmonics into the variation of the flux-density in the
teeth and yokes. This is the main cause of increased eddycurrent loss in
the core. The peak flux-density is also increased, and this contributes to
an increase in the hysteresis loss. It is difficult to estimate the magnitude
of these effects by simple formulations, but the finite element method
quickly brings the effects to light and provides an accurate means for
estimating them with a high degree of confidence.
12-27
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

Radius: 25,3, center: (0.0,0.0)


___ _ V a lu e s g1 BMQD
___ Values of BMQD

Fig. 12.19 Airgap flux-density variation under one pole with strong demagnetizing
armature reaction. Dotted line : magnets alone. Solid line ; resultant.

References
1. Fouad FA, Nehl TW and Dcraerdash NA [1981] Magneticfield modelling ofpermanent
magnet type electronically operated synchronous machine* using finite elementi, IEEE
Transaction], Vol. PAS-100, No.9, pp. 4125-4135
2. Reece ABJ : Electrical machine* and electromagnetics - computer aids to design, CJEC
Review, Vol.5, No.l, 1989, 34-41.
3. Hamdi ES, Licariao-Nogueira AF and Silvester PP [1993] Torque computation
mean and difference potential':, IEE Proceedings-A, Vol.140, No,2, 151-154.
4. Lowther DA and Silvester PP [1986] Computer-aided detign in magnetics, Springer-
Verlag.
5. Opera-2d Reference and User Guide, Vector Heidi Ltd., Oxford
6. Fitzgerald AE, Kingsley G and Umam SD [1990] Electric Machinery, fifth edition,
McGraw-Hill

12-28
13. EXAMPLES CALCULATED BY HAND
13,1 Introduction
The use and proliferation of brushless permanent-magnet motors over
the past ten years or so has been largely application-specific without any
standardization. For example, the motors used in computer hard disc
storage drives have been designed to keep pace with developments in the
disc drive. In a short period the data storage capacity of Winchester disc
drives increased 100-fold: the access time decreased by a factor of 10, and
tht! drive envelope decreased by a factor or 3 or 4. Needless to say, the
bmshless drive motors used in these drives have evolved at a comparable
Tiiie-
Another example of an application-specific brushless motor family is the
higli performance servo motor used to drive specific loads without the
use of gearing or clutches and brakes in order to achieve elegant
programmable motion. The configuration of these designs is driven by
high torque to inertia and ease of cooling the stator. Based on a casual
inspection of most vendor offerings it would appear that there are some
similarities, but their differences are more significant. It would seem that
among the various designs of similar motors each designer is able to
achieve acceptable performance for a particular physical size using
different magnet grades, number of poles and stator coils. There is also
a wide variety of feedback sensors in use. The author assumes that these
differences in design details from one vendor to another are attributed
principally to economic issues relative to each manufacturing operation.
The other possible reason would be that each designer achieves
acceptable results using design parameters which are within his or her
scope of understanding and experience.
Other motor products such as AG induction motors seem to all be alike.
If you disassemble them it is difficult to discern one from the other. DC
brush type motors are somewhat the same in their configuration. Most
companies which have developed brushless servo motors have been
known for their DC motors not AC motors. The large US induction
motor manufacturers were the last ones to enter the brushless AC or DC
servo business with new products so their design approaches are
somewhat different from those of the DC motor companies.
Most recently there have been several significant developments of
13-1
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

brushless motor by large AC induction motor companies. The strategy


employed involves using existing AC induction motor laminations,
housings, end bells and shaft/bearing systems. This low risk venture into
brushless motors requires only that a new permanent-magnet rotor be
designed to replace the aluminum die-cast squirrel-cage induction rotor.
The AC production machinery is easily adapted to wind and insert the
correct three phase coils in the stators. The use of existing tooling and
assembly equipment provides an immediate cost-effective brushless motor
potential with very little investment.
This chapter provides examples of two brushless motor designs. The first
is an interior-rotor configuration similar to the example of a converted
AC induction motor. The second exemplifies the exterior-rotor brushless
motors commonly used in disc drives or fans and blowers. The examples
illustrate different sets of design constraints which dictate the use of
different magnet grades.
The brushless motor designed from the AC induction motor would be
limited to adjustable speed. Its torque/inertia ratio would seem to be
quite high because the rotor is inside the stator. However, the only
possible permanent magnet which could be used is a ferrite grade
because of the narrow cross-section of the stator teeth, and in some
instances the yoke cross-section. It is possible to use a bonded rare-earth
magnet in many instances, but the cost would probably not be justifiable.
The low flux/pole, requiring a large number of turns per coil, would
lead to a high inductance which is undesirable in a good servo motor.

13.2 Interior-rotor motor designed from an AC induction motor


Fig. 13.1 shows the cross-section of a typical single-phase induction motor
with 24 stator slots. The objective is to use the unmodified stator
punching with the production stack lengths using the tooled slot liners
(with cuffed ends to eliminate end insulators) and slot wedges. The
brushless version of the stator must be manufactured on the existing
stator line using all tooling and machine stations, with the only
modification being the reprogramming of the winders to produce three-
phase windings rather than the two sets of windings required for the
single-phase AC motor. The ball-bearing system, end-bells and housing
would also be used without modification, allowing the final assembly
operation to be utilized unchanged.
13-2
13. E x a m pl es c a l c u l a t e d by h a n d

Fig. 13-1 Cross-section of DC brushless motor derived from AC


induction motor
The first task is to identify the parts of the AC induction motor which
must be used to design the magnetic circuit of the brushless motor and
to list the dimensions. The relevant dimensions of the AC induction
motor stator (in inches) are listed in Table 13.1:
Lamination OD D. 5.000
Diameter over slots 4.140
Yoke thickness {D, -D ) /2 - V 0.430
Lamination ID D 3.000
No. of slots 24
Tooth width *r 0.230
Slot opening % 0.080
Lamination material M-Sfi
Lamination gauge #24 0.025
Core loss at 60 Hz, lOkG 1.10 W/lb
Core loss at 60Hz, 15kG 2.40 W/lb
Stack length ^dc 1.25/2.50
T a b le 13.1 St a t o r sta c k d im e n s io n s

13-3
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

The laminations, stack lengths, stator ID and winding area are fixed.
Since the brushless motor must be low in cost to be competitive in the
adjustable-speed market, the use of the highest grade Ferrite magnet in
terms of remanence must be employed. The coercivity is less important
because the low-cost magnet is best utilized with extra thickness to
prevent demagnetization. The inverter drive will provide ultimate
protection because the drive transistors will be protected with a current-
limit circuit which also protects the magnets from demagnetization. The
highest possible flux/pole is then the main objective.
The Allen-Bradley/TDK material known as FB4B is easily selected for its
properties given in Table 13.2:
4000 G
3200 Oc
c, 3300 Oe
Temp. coeiTt. of Br -o.i8 % r c
T able 13.2 P roperties o f A l l e n -Bradley /T D K fb4b m agnet
A B /H curve from the vendor catalogue is shown in Fig. 13.2 complete
with the load-line due to the magnetic circuit airgap and the load-line
resulting from the stator current
The general practice of using mechanical airgaps of 0.030" nominal with
Ferrite permanent magnets in DC commutator motors is applied to this
brushless motor design. Since the speed is not intended to exceed 4000
rev/min, the only magnet retention will be a 3-M structural adhesive
between the rotor lamination stack and the magnet arc inside radius.
To achieve a maximum flux at +60 C rotor operating temperature, and
to prevent demagnetization at the peak current at starting, an
approximate permeance coefficient B/H of 10 is selected (see Chapter
4). This translates to a magnet thickness of 0.300":

B IH - iM - 10;
g (13.1)
= 10 * 0.030 = 0.300in.

13-4
13. E x a m pl e s c a l c u l a t e d By h a n d

]GBj M

B/H curve of Allen-Bradley/TDK magnet material FB4B


<

Pig- 13.2

13-5
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

To allow for sintering and grinding tolerances the actual magnet


thickness will be specified as 0.310" with the actual tolerance over 80%
of its area in accordance with Allen-Bradley/TDK policy. This means that
for magnetic design 0.300" thickness is used, but in mechanical
dimensions a nominal thickness of 0.310" is used. If higher speeds are
required and magnet retention is required, such as a shrink-ring of 303
stainless steel or wrappings of Kevlar yarn and epoxy, the appropriate
increase in air gap would be used. The magnet thickness should then be
increased somewhat, but there is a practical limit on the thickness
relative to the rotor diameter. The reason for this is that the pole area
is calculated using the inside radius of the magnets rather than the
outside radius or a mean radius. Remember the flux density jBm from the
B-H curve is higher at the inside radius of the magnet and lower at the
outside radius. The easiest way to rationalize this is to convert flux-density
to lines of flux $ = B x A, where A = inside pole area. The outside pole
area at the air gap is larger, but the total number of lines $ with radial
orientation is identical. Therefore, the flux density on the pole OD is
lower than on the inside.
The next issue concerns the selection of the pole number to be used
with the three-phase 24-slot stator. If reference is made to Table 3.4 the
possible numbers of poles with 24 slots are 2, 4, 8, 10, 16 and 20. The
winding pitch (for a lap winding) must be practical and within
reasonable limits so that the coil end turns are not too large. Since this
design will be wound on existing AC induction motor equipment, a
certain length of end turn is required for automatic insertion. Therefore,
the winding pitch should be minimal, which implies a larger number of
poles on the rotor.
The commutation frequency for 6-step operation corresponding to 4000
rev/m in can be determined by the equation
4 mm = x rev/sec * pole-pairs
6

r rev/min poles
= 6 60 2 (13.2)
rev/min * poles
20

13-6
13. E x a m pl es c a l c u l a t e d by h a n d

I pr winding pitch is determined by dividing the number of poles into


the number of slots and rounding down to the next lowest whole
dumber.
No. of poles 2 4 8 10 16 20
Slots/pole 12 6 3 2.4 1.5 1.25
Winding pitch 12 6 3 2 1 1
/omm at 4000 rev/min 400 800 1600 2000 3200 4000
[Hz]
Commutation interval 2.5 1.25 0.625 0.5 0.313 0,25
[ms]
Tablf, 13.3 S elec tio n o f po le nuniber
The best design would probably be the 10-pole because the winding pitch
is small and the inherent detent or cogging torque is small due to the
fad that only two out of 10 poles align with stator slot-openings as
rotation occurs. The commutation frequency is high for the M-36
lamination grade, due to corc losses which would need to be analyzed if
a 10 pole-design were chosen.
The requirement to use existing winding and insertion equipment
restricts the selection to 2, 4, or 8 poles, each of which is an integral-slot
design. As the rotor rotates, each pole edge will align with each slot in
the stator causing a certain amount of cogging torque. Such a design
would have difficulty in the marketplace. The tooth tips of the existing
laminations are not so thin as to become saturated from the oncoming
edge of a rotor pole as rotation takes place causing extra detent torque
luch as could be the case with certain AC induction motor laminations.
Several ways to reduce the detent or cogging torque are listed in section
17, Table 4.3. Since this design must use existing stator stacks which are
itraight and not skewed, and the rotor magnets must be tooled from
scratch, the best solution to the cogging would involve the permanent
magnets. They could be either sintered with chamfers or radii at the pole
edges, or if full-pitch arcs are used the poles can be skewed during
magnetization.
The 2-pole option is not practical because the end-turns would be
13-7
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

excessive due to the 12-slot winding pitch. The choice could be either a
4-polc rotor or an 8-pole rotor. In either case a double-layer lap winding
could be used with 24 coils or 8 coils per phase. The lap winding
requires hand insertion. The most practical choice for automatic
insertion is a single-layer winding pattern of 12 coils with 4 coils per
phase, i.e. a concentric consequent-pole winding. If this pattern were
used with a four pole rotor, there would be 2 concentric coils per pole
per phase, which would probably require coil insertions of a single phase
at a time. With the use of the &-pole rotor, the concentric winding has
only one coil per pole per phase, permitting automatic insertion of all 12
coils into the stator in one stroke of the machine, which certainly speeds
up production.
So far the design parameters can be summarized as follows:

No. of stator slots 24


No. of rotor poles 8
No. of stator coils 12
No. of coils/phase 4
Winding configuration Concentric
Consequent-pole
Winding pitch 3 slots
T a b le 13.4 Sum m ary o f d e sig n d e ta ils

Next to be determined are the pole arc and winding connections, wye or
delta. Reference can be made to the generic back-EMF wave shapes For
3 slots/pole, shown in Fig. 3.15a and b. The controller will either be a 6-
step squarewave drive energizing only two phases at a time, or a sinewave
drive with current in all three phases simultaneously. The 6-step drive is
much lower in cost for this size of motor than the sinewave drive and
probably more suitable for this product for the market it would serve. If
that is the case, the selection of a delta connection would be appropriate
with a magnet pole-arc of 2/5 of one pole-pitch or 30. If a full-pitch
magnet arc of 45 is chosen the wye connection should be used,
otherwise the third-harmonic circulating in the delta would be excessive.
13-8
13. E x a m ples c a l c u l a t e d by h a n d

C ontinuing the reference to Fig. 3.15a and b, if a sinewave drive is used


tficwye connection would still need to be used with 45 magnet arcs, but
either wye or delta could be used with 30 magnet arcs. In both cases the
sinewave back-EMF quality is not particularly good even with skew.
Sauce this motor will be driven by a 6-step drive and the magnet pole arcs
dill be full-pitch or close to 45, the best choice is the wye connection,
the reason for the full-pitch magnet arc is for low cost and manufactur
ing simplicity. The magnets will be designed as 90 arcs with a negative
tolerance. With four magnets epoxied to the rotor lamination stack, a
magnetizing fixture can be made which will magnetize a pole pair on
each arc simultaneously after the magnets are attached. Very accurate
poles will result (as good as the magnetizing fixture) which can be easily
jltcwed to eliminate cogging. The actual arcs do not require skewing,
only the magnetized poles. The rotor could be fixed in the magnetizing
head to set the poles within each physical arc such that there is one full
pole in the centre of the arc and two half poles, one on either side. It
would not seem to make much difference how the magnetization is
oriented as long as the actual spaces between the 90 magnet arcs are
tept to a minimum.
Rotor dimensions are summarized in Table 13.5. (Magnet adhesive is
ignored because its approximate 0.005" thickness is statistically accounted
for in the maximum magnet thickness and the nominal airgap.)
Rotor OD Dr =D-2g 3.00 - 2 x0.090 = 2.940"
Rotor yoke diam. O ^ D r -iL u 2.940 - 2 x 0.310 = 2.320
Magnet thickness *14 0.310 max.
Airgap length e 0.030 nom.
Physical magnet arc 90 max.
Magnet arci per 4
motor
I aiilk 13.5 Summary o f r o to r design details
Thr calculation of the useful flux per pole 4> is the next step. The two
Hack lengths in the design are 1.25" and 2.50. Since the pole-arcs are
full-pitch, a magnet overhang at each end of the stack any longer than
13-9
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

half of the magnet thickness would be wasteful, due to leakage from pole
to pole at the overhang. The resulting rotor lengths are determined by
adding the magnet thickness to the stack length
Short rotor length Lrl = 1.25 + 0.31 = 1.56 in. (min);
(13.3)
Long rotor length= 2.50 + 0.31 = 2.81 in. (min).
The magnet pole area is calculated as follows:

poles (13.4)
n *_2.3j * 1.56 _ j 42 in 2 for short stack
8
and
Anp = n g = 2.56 in 2 for long stack. (13.5)
With the magnet flux-density at 60 C equal to 3300 G (Fig. 13.2),
$ = BmAm * 2.542
m m (13.6)
= 3300 * 1.42 * 2.542 = 30,234 lines for short stack
and
$ = 3300 x 2.56 * 2.542 = 54,506 lines for long stack. (13.7)

In Wb, = 3.023 x 10 "4 for the short stack and 5.451 x 10 ~4 for the
long stack. The magnet poles should be tooled using the minimum
number of lengths to provide the two stack lengths. One possibility for
this motor would be two lengths, 0.70" long and 0.86" long. Sixteen arcs
0.70" would yield the 2.81" rotor magnet and four 0.70" arcs plus four
0.86 long arcs would yield the 1.56" long rotor magnet in a staggered
configuration so that the slot openings between them do not line up.
The spaces caused by the plus zero to minus 0.5 tolerances on the 90
arc angle would cancel.

13-10
13. E x a m pl es c a l c u ij \ t e d by h a n d

The other possibility is to tool the correct length for each stack length
of 1.42" and 2.56" to eliminate the extra labour of the rotor assembly
when handling so many arcs. Even though the number of arcs to be
tooled is the same as for the shorter lengths, the longer arcs do not offer
the flexibility of adding other rotor lengths without tooling another arc.
The next step is to estimate of the number of turns per coil. One way to
do this is via the EMF constant Sincc the maximum operating speed
is to be 4000 rev/min, a good rule of thumb for a Ferrite motor is to
begin with the rated speed at 80% of the no-load speed, giving

WNL = = 5,000 rev/min. (13.8)


0.8

and
<*>nl = ^ = 524 rad/s. (13.9)
(For rare-earth magnets 90% would be a good place to start.) The
inverter will be powered from 115V AC single phase, full-wave rectified
to provide a 160V DC bus. A rough calculation of the EMF constant
and torque constant kj. can be determined by setting the back-EMF at the
no-load speed equal to the DC bus voltage, neglecting losses. Thus

*E = M = 0.30 V-s/rad (13.10)


and
kT = 0.30 Nm/A = 2.66 lbf-in/A. (13.11)
The last point to be made before the turns are calculated is the
difference between peak and kj. If the back-EMF is trapezoidal with
overlap from phase to phase the peak equals the average. Referring to
Fig. 3.156, the back EMF waveform with wye connection and one full slot
skew calculated for the three slot/pole family indicates a 100%
commutation average. In practice due to the airgap and slot openings
the winding distribution will not be quite that good so we can be safe by
estimating 90% commutation zone average, i.e. allowing a factor C = 0.9
for winding and flux distribution effects. Equation (3.3) can now be used
to calculate the total number of conductors 2 in the machine. Assuming
o = l parallel paths,
13-11
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

Z = - x x----- V - '-30---------------------------- = 1299.(1


2 0.9 3.023 * 10-4 * 4
Since there are 12 coils each with 2 sides, the number of turns/coil must
be 1299/(2 x 12) = 54 turns/coil. Since this winding is a single layer type
which means that only one coil side is inserted into each slot, a
reasonable fill factor is 60% of the available area after accounting for the
accounts for the slot liners and wedges (see equation (3.14)).
The slot area can be determined by calculating the swept area of the
entire slot region and subtracting the teeth area, then dividing the result
by the number of slots.
Swept area = - Z?2]
4 (13.13)
= -^[4.142 - 3.002] = 6.39 in2.

[ L - D\1vVsloLs
Tooth area = wT-----------

= 0.230 * I f-I 4 ~ 3 I- x
(13.14)
-------------- * 24
Z'f
2
= 3.15 in2.

Slot. area = -----


Swept area - Tooth area
---------------------------
Alerts (13.15)
6.39 - 3.15 0.13 in2.
24
The maximum wire diameter which will fit is determined by assuming the
wire is square (known as square bedding the wire for extra winding
space) and solving for the wire diameter Z)w using equation (3.14):
-^siot ^sipt
N N (13.16)
0.13 x 0.6 = 0.038 in.
54
13-12
13. E x a m pl es c a l c u l a t e d by h a n d

The nearest wire gauge is AWG #19 with a maximum diameter over
insulation of .0386, a bare wire diameter of 0.0359", a resistance at 20C
of 8.0293/1000ft, and a weight of 64.04 oz/lOOOft.
The winding resistance can be calculated after the mean length of turn
(MLT) is determined as twice the slot pitch to the average centre of the
slot opening plus twice the stack length plus two times the allowance for
end turn height on the automatic insertion machines: for the short
machine
MLT = 2 x (1.25 + 0.375) + 2 * 1.785 = 6.82 in. (13.17)
For the long machine MLT = 9.32". The resistance per coil is therefore

R . = i 4 * 6 82 x 8.029 = 0.246 Cl (13.18)


0011 12 x 1000
and with 4 coils per phase the phase resistance is 4 x 0.246 = 0.986 3.
The line-line resistance with 2 phases on and wye connection is 2 x 0.986
- 1.97 fl at 25 C.
At 4,000 rev/min the back-EMF is = 0.30 x 4,000 x rc/30 = 126 V.
The supply voltage is 160 V, so that with an allowance of 2V for transistor
volt-drop and an effective supply resistance of 1 Q, the current supplied
from an unregulated controller would be (160 - 126 - 2)/(1.97 + 1) =
10.7 A. This gives rise to a torque of 10.7 x 2.66 = 29 lbf-in. The
corresponding power is 29/8.85 x 4,000 x 7t/30 = 1,372 W = 1.84 hp.
The current density at maximum power should be checked to determine
approximate continuous rating. With 120 squarewave currents of peak
value 10.7 A in each phase the RMS current is 10.7 x /(2 /3 ) = 8.74 A.
Dividing this by the cross section area of the copper conductor,

j _ ^RMS
Bare wire area
8 74 (13J9)
------ ------ = 8,634 A/in2.
-4 x 0.03592

13-13
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

The maximum current-density allowed for a motor of this type without


cooling would be about 3,500 A /in2. With continuous air-over cooling
about 7000 A /in 2 might be appropriate. Therefore, as an initial estimate
of the design before actual samples are built and thermal resistance tests
are completed to finalize the ratings, a reasonably good estimate of the
rating can be made based upon the current-density calculation at 4000
rev/min using a current density of 3500 A /in 2 or about 4.4 A, which
gives a torque of 11.7 lbf-in and a power output at 4,000 rev/min of 554
W or 0.74 hp. With a phase resistance of 0.986 and two-phase-on
operation, the copper losses are 2 x 0.986 x 4.42 = 38 W. With air-over
cooling and a current-density of 7,000 A /in2, the rated current is 8.8 A,
which gives a torque of 23.4 lbf-in, a power output of 1107 W (1.48 hp)
at 4,000 rev/min, and copper losses of 153 W. Fig. 13.3 shows the
speed/torque curve for the 1.25" long motor including the continuous
operating range for both cooled and uncooled operation.
A very rough estimate of the core losses can be made as follows. Assume
that at 60 Hz half the core losses are attributable to hysteresis losses
proportional to and the other half attributable to eddy-current
losses proportional to if2/ 2. The relevant frequency at 4,000 rev/min is
4,000/60 x 4 = 266 Hz. If the flux-density in the steel is taken to be
roughly twice the flux-density in the airgap, i.e. approximately 0.6T, then

Continuous operation at 4.4 A


^Intermittent operation at 8.0 A
M ax current & torque

Fig. I3.S Speed/torquc curve of DC brushlcu motor derived from AC


induction motor
13-14
13. E x a m pl e s c a l c u l a t e d by h a n d

the core losses per lb can be estimated by scaling as

t = i i X ( M f X 266 + l i , ( M Y X f M 6 ?
R 2 U-Oj 60 2 ^ 1.0 J { 60 ) (13.20)
= 1.08 + 3.89 = 5.0 W/lb.
From the stator dimensions calculated earlier the iron volume is
approximately 16 in3 and if the density is 0.28 lb/in 8 the core weight is
16 x 0.28 = 4.6 lb. Therefore the corc losses are approximately 5.0 x 4.6
a 23 W- If bearing and friction losses at 5 W are added, the total losses
at 4-4 A are 38 + 23 + 5 = 66 W, for an efficiency of 554/ (554 + 66) x
jftf) = 89.3%. At the 8.8 A operating point the total losses are 153 + 23
f 5 = 181 W and the efficiency is 1107/(1107 + 181) x 100 = 86.0%. An
obvious improvement would be to build the motor with thinner
laminations or to use a higher grade of core steel, but these efficiency
figures are considerably better than what could be achieved by the
induction motor in the same lamination, which would probably have
twice the losses at a comparable operating point
With no current-limit in the inverter the locked-rotor current could be
as high as 160/1.97 = 81 A, assuming no impedance or voltage-drop in
the supply.
In Fig. 13.2, a load-line was drawn on the vendor B /H curve for the FB4B
rotor magnets at the maximum rotor operating temperature of 60C.
The maximum allowable demagnetization field Hmax can be taken from
the plot at 20 C as -3200 Oersteds. The demagnetization current with
wye connection can be calculated from equation (3.6) using this Hmax
value, assuming a = 1 parallel path through the stator. With two phases
on, the number of conductors carrying current is z = 54 turns/coil x 4
coils/phase x 2 conductors/turn x 2 phases = 864: thus
1000 x 4 x l x 4 x (0.30Q 4 0.Q30) x 320Q

tox 4w * 39.37 864 (13.21)


= 39.5 A.
The locked-rotor current exceeds the demagnetization current by 2:1, but
both these currents are far in excess of the rated load current, for which

13-15
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

the transistors will be rated. Current-limit in the inverter is therefore


essential to protect the transistors, probably at a level no higher than
12-15A, and unless grossly oversized transistors are used, the risk of
demagnetization is eliminated by the current-limit
With such a value of current-limit the magnet thickness could he
reduced, possibly to as litde as 0.20". This would be acceptable if the
motor was thermally protected and it would still be possible to provide
a starting torque of 15 A x 2.7 lbf-in or 40 lbf-in, about 4 times the rated
torque.
The final step would be to build one or two samples and subject them to
a rigorous design verification procedure by testing as described in
Chapter 11. Minor adjustments in the number of turns and/or wire
gauge might be required. Subsequendy the 2.5" long motor (or any other
length) can be easily calculated using parts of the foregoing procedures
along with test data gained from the 1.25" long motor.

13.3 Exterior rotor disc drive m otor design


Computer disc storage drives contain two motors. One is a limited-
rotation motor known as a "voice-coil actuator" which is also a type of
brushless permanent-magnet motor. It has a fixed magnet and a moving
coil attached to the arm on which the read/write is mounted, giving a
very low inertia. The coil is connected to a control circuit with flexible
lead wires.
The other motor used in a disc drive is the hub or platen drive. Floppy
disc drives usually use low-speed axial-gap motors, but the fixed disc or
"hard" disc usually uses an exterior-rotor motor whose rotor carries the
discs themselves. The only practical configuration for these motors is to
have the permanent magnet rotor mounted inside the aluminum disc
hub. The stator assembly is then attached to a shaft or mounting base.
These motors often use remote- sensing integrated circuits for
commutation, so that there is no need for Hall sensors in the motor.
This saves space and five extra electrical connections.
Fig. 13.4 shows a typical cross-section of a "motor-in-hub". There are
many variations of this basic configuration and the analysis given in this

13-16
13. E x a m pl es c a l c u l a t e d by h a n d

Fig. 13.4 Motor-in-hub configuration for computer fixed disc drive


section is limited to the magnetic and electrical design. The specialized
mechanical engineering issues are not discussed here. However, the
requirements listed in table 13.6 are worth noting as they influence the
magnetic design.
Wllh limited space available, there can be no wasting of space. The
balance of copper, iron and magnets must be optimized with the
minimum of unused air space. The following design decisions are nearly
self-evident and can hardly be challenged.

13-17
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

Motor-in-hub requirement*
Exterior-rotor motor
Very little space available
Minimum cost
Long life
Low noise
Constant ambient temperature
Low power density
T able 13.6 G enera l r e q u ir e m e n t s o f c o m pu t e r d isc drive

Magnet grade Bonded or sintered Rare-earth


No. of phases 3 with bipolar drive
No. of sensors None. Remote sensing.
No. of poles 6,8, or 12 (possibly 10 with full-pitch winding)
No. of slots 6 or 9 (sometimes 12 or 15)
T able 13.7 I n it ia l desig n d e c isio n s f o r d isc -drive m o t o r

These motors produce very litde starting or running torque which means
that the cogging or detent torque must be kept to an extremely low level.
This will also help to reduce the audible noise. To accomplish this, the
rotor poles or the stator should be skewed. In addition, the stator teeth
tips should be as thick at their root as they are long, extending from the
tooth, to prevent tooth tip saturation.
This also reduces cogging and audible noise. The stator slot openings
should be very small, just enough for the magnet wire to enter. Ihe use
of needle winders for interior stators is not recommended because too
much air space is required for the needle. 'Ihe winder of choice is known
as a "fly winder" with winding jaws or guides to feed the wire into the slot
openings.

13-18
13. E x a m ples c a l c u l a t e d by h a n d

DC supply voltage 12 V nominal, 9,6 V min


Speed 3fi00 rev/min
Maximum diameter 0.875" (22.2 mm)
Maximum length 0.400" (10.2 mm)
Stator ID (shaft) diameter 0.197" (5.0 nun)
Back-EMF constant *E 1.5 V/krpm
Torque at full speed 1 oz-in
Table 13.8 S pecification data

A first attempt at design will be based on one or two more assumptions


and then refined by reiteration, if required. First, the airgap between the
rotor and stator will be kept as small as possible without grinding
requirements, i.e. 0.010" on the radius (or less). The magnet thickness
will be the next consideration so that a cross-section sketch can be made
to design a lamination. In the last problem the magnet thickness was
chosen to yield a high value of flux 4> but in this example there is not
enough space so a different load line is used. The current is very low in
this application, so demagnetization will not be a concern. A good first
choice for an 8-pole rotor is a magnet thickness of 0.040". For a magnet
material, a bonded NdFeB will be considered first because of its excellent
performance in relation to its cost.
Temperature is seldom a problem with these motors because of the
controlled environment. The low power and losses in the motor keep the
temperature very low so that NdFeB works very well. Fig. 13.5 shows a
B/H curve from the vendor, with the operating load line plotted for an
airgap of 0.010" and a magnet thickness of 0.040". This yields a flux
density ^ of 4,900 Gauss when projected to the vertical axis of the 75 C
B/H curve. This value is based on a bonded 360 ring magnet made from
MAGNEQUENCH MQ1 powder which is one of the lowest cost
permanent magnet grades available for its flux output. It would be
supplied with a coating of epoxy paint as required by the disc drive
industry to prevent magnetic contamination in the clean rooms used for
assembly.

13-19
D e s ic n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o to rs

1.0 B/H = 4
14
B (kG)
12

10
8

6
4.9 kG
4

Fig, 13.5 B/H curve of Magncqucnch magnets

The dimensions of Fig. 13.6 can now be determined by calculating the


cross-section required of the rotor yoke. With the rotor OD given as
0.875", airgap 0.010" and magnet thickncss of 0.040", the lamination and
rotor can be dimensioned. For example, a first estimate of the rotor yoke
would be about 0.050" for an 8-pole full-pitch magnet.
Rotor OD Dro 0.875"
Magnet OD 0.875 - 2 x 0.050 - 0.775"
Rotor ID 0.775 - 2 x 0.040 - 0.695"
T able 13.9 initial dim ensioning
The flux/pole can be calculated by first determining the pole area on
the inside diameter of the magnet with a maximum rotor length L^ of
0.370" and an 8-pole magnet: with a 2 allowance between magnet arcs,

13-20
13. E x a m pl es c a l c u l a t e d by h a n d

= ir/feA x 4 y
p 8 45'
ir x 0.695 x 0.370 wX 43 (13.22)
8 45
= 0.097 in2.
Consequently the flux/pole is
> - ^ A p *2.54 2
(13.23)
= 4900 x 0.097 x 2.54* = 3066 lines

or 30.7|iWb/pole. The next step is to check the flux density in the 0.050"
thick rotor yoke. The flux 4> for each of the 8 poles is split into two
paths in the rotor yoke, giving
B = ---------- 59^6/2---------- = 12 g4 kG (13.24)
v 0.050 x 0.370 x 2.542

or 1.57 T. This value is acceptable as the soft iron of the magnet return
path material is saturated at 18 kG using low-carbon steel. If a six-pole
magnet ring were used, the flux/pole would be 33% higher but the same
yoke thickness could be used. A summary of rotor dimensions which have
been determined so far is given below.

Rotor OD Dro 0.875


Rotor ID 0.695
Airgap e 0.010
Rotor length 0.370
No. or poles 2p 8
T a b le 13.10 F u r t h e r r o t o r d im e n s io n s

13-21
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

The stator outside diameter equals the rotor inside diameter minus twice
the airgap:
Ds Dn ' (13.25)
= 0.695 - 0.020 = 0.675 in
or 17.1 mm. For an 8-pole motor there are very few practical choices for
the number of slots for such a small compact design as is being
attempted here with the rotor on the outside. 6, 9 or 12 are probably the
only possibilides. The 6-slot stator would be easy to wind, but with so few
coils the end turns would be too large to fit into the space. The 12-slot
stator would work very well in a large enough diameter motor to have
enough circumference on the stator lamination for 12 slot openings.
The best selection might well be the 9-slot stator with 8 rotor poles in
terms of end-turn height and slot openings. The other significant
advantage is that only 1 out of 8 pole edges would line up with a slot
opening at any time during rotation. This results in a low cogging
torque. The disadvantage of the 8-pole, 9-slot design is that both known
winding patterns can cause unbalanced magnetic radial loading which is
said to cause noise problems, (cf. Fig. 3.12). If that is the case, perhaps
a 6-pole rotor should be considered with a 9-slot rotor. However, if space
permits the 8-pole 12-slot design would yield a good motor with low
cogging and quiet running.
Continuing with the 8-pole, 9-slot design, the lamination stack length
must be selected before the tooth thickness can be determined. The rule
of experience used from the interior-rotor example was that the magnet
should overhang the stack by about 1/2 its thickness for full-pitch poles.
The magnet length in the axial direction is 0.370" and the thickness is
0.040" so the stack length Lstk should be about 0.330 max. If .018" thick
laminations are used of M-19 annealed stock, about 18 laminations would
be required, which would figure out to 0.324" nominal, plus or minus
some tolerance. It is possible that 0.014" thick laminations should be
used for lower losses.
With an allowance of 18 kG in the stator teeth, and based on the
assumption that each of the 9 teeth will collect the flux from a single
pole, the tooth width can be determined as follows, allowing for 3
leakage factor (gap flux/magnet flux) of 0.9:

13-22
13. E x a m pl es c a l c u l a t e d by h a n d

9 BjL<k * 2.S42
* (13.26)
0 9 * 8 * 3066' 9 = 0.0652 in.
18,000 0.324 x 2.542
The lamination cross section can be drawn to scale, preferably on a GAD
system so that the winding area can be determined by the computer. The
slot openings are maintained at 0.040" 0.001" with a careful tooth tip
shape using tapered teeth and a 0.030" radius in the comers of the tip
so that tip saturation will not cause cogging.
The CAD system reveals a total slot area of 0.0132 in 2 or 8.52 mm2,
allowing for 0.005" build-up of epoxy insulation, i.e. 0.0066 in 2 for each
coilside in a double-layer winding. The mean turn length is easily
calculated as before using twice the stack length plus twice the tooth
thickness plus insulation and four times the coil thickness which is
estimated as 0.04": thus

MLT = 2 * (0.324 + 0.010) + 2 * (0.10 + 0.01) + 4 * 0.04 (13.27)


= 1.05 in.
The stator yoke thickness needs to be about the same as the rotor yoke
thickness except that it should be slightly thicker by the ratio of the rotor
length to the stator length for the same flux density. This works out to
be 0.050 x0.375/0,324 = 0.058". If the shaft diameter is 5 mm (0.197")
then the diameter at the bottom of the slot is 2 x 0.058 + 0.197 = 0.313".
A full radius is shown in the slot bottom in Fig. 13.6.
This completes the lamination design, except the gauge which could be
either 0.014" or 0.018" thick M-19 silicon steel, annealed after punching.
It might be more cost-effective to use M-15 silicon steel fully processed
before punching rather than M-19. The benefit and final decision based
on performance (core loss heating) vs. cost can only be determined
precisely by obtaining accurate quotation on the two materials and
building laser cut samples of each one and taking test data.

13-23
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

Fig. 13.6 Motor-in-hub crojj-seclion


Certain lamination vendors claim that the cost of annealing after
punching is the same or lower in cost than not annealing. This because
they have made it part of their standard automated process. If such a
vendor is verified, then it would probably be most cost effective to use
their core loss equivalent to M-19 annealed in their automatic process
which also would produce an iron oxide insulating film on each
lamination surface.
The stacking of the laminations must be done carefully with all of the
punching burrs going in the same direction. For such a small motor
stack, there should be no burr reversals in a stack or the iron content will
13-24
13. E x a m pl es c a l c u l a t e d by h a n d

vary from motor to motor. The method of holding the stack can be best
implemented by one of three ways listed below:
Press laminations on to a knurled shaft; epoxy coat; and wind
Clamp the lamination slack in alignment fixture; epoxy spray coat; wind;
then press on to shaft
Have the stamping die fabricated to provide cleatcd or dimpled stacks from the
punching operation; Insulate and wind
T able 1 S.l 1 M eth o d s o f fabricating stator
The actual choice depends on the production system of the motor
vendor plus the details of the mechanical design of the motor such as
how the starts and finishes of the phase windings will emerge from the
wound stack, under the ball bearings and out to the driver board.
The next step is to determine the number of turns per coil and the wire
gauge which will fit into this small lamination slot with a terminal
resistance acceptable to the design. With reference to Fig. 3.12b, a delta
connection with full-pitch magnet arcs would have excessive third-
harmonic in the phase EMF and this would cause circulating current in
the delta. This could be avoided by magnetizing the magnets in such a
way as to get 120 magnet arcs. Alternatively a wye connection could be
used. If the wye connection is used with full magnet arcs and one slot-
pitch of skew, the Fourier series analysis shows that the commutation
zone average back-EMF is 65.6% of the formula value. The specification
for is 1.5 V/krpm = 0.0143 V-s/rad so that Aj- = 0.0143 Nm/A or 2.03
oz-in/A. Using equation (3.3) or (7.28), and again using the leakage
factor (gap flux/magnet flux) of 0.9,

7 - x k*n x 1 x 1
3
2 <bpy 0.656 0.9 ......
(13.28)
= 2 x 0-0143 x 7r x 1 x _ 1_ = 93Q
2 3,066 *10' x 4 0.656 0.9

conductors, giving 930/(2 x 9) = 52 tums/coil. The winding space per


coilside was determined by the CAD layout of the lamination to be

13-25
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

0.0066 in 2 and the MLT to be 1.05". Therefore, the winding gauge and
the terminal resistance can be determined using the same method as in
the previous example:
^slot 'slot
\ N (13.29)
0.0132 x 0.6 0.0087 in.
2 * 52
No. 32 AWG wire is 0.094 over insulation with a bare wire diameter of
0.0080" nominal. The resistance per 1000 ft. is 162.0 Q at 20 C and its
weight is 3.28 oz/1000 ft. The resistance per coil is determined as follows:

/? = 5?_ 105 x 162.0 = 0.737 fl (13.30)


C01 12 * 1000

With three coils per phase the phase resistance is 3 x 0.737 = 2.2 Q and
with wye connection the line-line resistance is 2 x 2.2 = 4.4 Q at 20C or
5.35 Q at 75C.
The specification called for 1 oz-in of torque at speed, requiring a
current of 1/2.03 = 0.5 A. At 75C winding temperature the copper
losses would be 0.5 2 x 5.35 = 1.34 W. The RMS current in each phase is
0.5 x 7(2/3) = 0.408 A, so the RMS current density is 0.408/(k /4 x
0.0082) = 8120 A /in2.

13.4 Summary
The basic process of manual calculation of a design is the same for all
brushless motor designs including servo motors. The requirement,
packaging, cost and production methods help the designer select the
magnet grade, interior vs. exterior rotor, number of slots, poles and
winding method. The performance goals are laid out, the flux per pole
is calculated, then the turns per coil and finally the wire gauge and line
to line resistance and the losses.
This process can be reproduced on the computer using either
spreadsheet formulations of the basic design formulas, or powerful

13-26
13. E x a m pl es c a l c u l a t e d by h a n d

special-purpose CAD programs like PC-BDC. (Chapter 12). No matter


how powerful and accurate the computer programs available, the
judgement of the design engineer is always required and is always final.
This judgement rests on the ability to calculate designs "on the back of
an envelope", and is no less important today than it was before computer
software became available.

13-27
14. CONTROL SYSTEMS PERFORMANCE
14.1 Introduction
Brushless permanent-magnet motors have a linear torque/current
characteristic, low torque ripple, and fast response. This makes them
highly suitable for controlling speed, position or torque, either in a
single-quadrant variable control mode, or as four-quadrant servomotors
with reverse speed capability and dynamic braking.
This chapter describes the brushless permanent-magnet motor and its
controller as a control system. It includes a review of the simplified
mathematical models that describe the motor and controller, followed by
the main types of control system. Control system design is described in
relation to closed-loop operation, frequency response, step response,
stability, steady-state error, root loci, lead/lag compensation, pole
placement, and robustness. The design and tuning of PID controllers is
discussed in both linear (analog) and digital forms. Z-transform methods
are described for discrete digital systems. The chapter concludes with a
review of modern control methods including adaptive control, optimal
control, and observers.

14.2 Basic modelling tools for linear control systems


14.2.1 Laplace transforms
For what is called linear control systems analysis, the motor/drive system
must be described mathematically, usually in the form of a model based
on its differential equations. These equations include the mechanical
equations of motion and the electrical circuit equations, together with
equations describing the control loops for torque, velocity, etc. The
mathematical model is used to predict and characterize the system
response to a variety of stimuli or "inputs". The response might be a
graph of position or velocity vs. time, and the "inputs" might be changes
in load or changes in control signals.
Manipulation and application of the system differential equations is made
easier by applying the Laplace transform. The formal definition of the
Laplace Transform is

14-1
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

F[s) = tf [/(*)] * Jo{"X Q e~ dt (14.1)


where J[t) is a function of time that represents either an input such as an
applied voltage waveform, or an output such as speed vs. time; the j,

Laplace operator, is the complex frequency. The formal definition of the


Laplace transform in equation (14.1) appears not to have any physical
significance, but this is because it is a mathematical tool used, in effect,
to convert the differential equations of the system into algebraic
equations, so that the system response can be studied using simple
algebraic and graph-plotting techniques instead of the differential
calculus. Implicit in this transformation is that the foWresponse function
J[t) is exchanged for a frequency-response function F(s), which is often the
most informative way to express the characteristics of the system.
After Laplace transformation, a differential equation in the time variable
t becomes an algebraic equation in the frequency s. In general, s is a
complex number a+jo). The imaginary part of the frequency, co (rad/s),
is what engineers normally think of as frequency in the sense of steady-
state sinusoidal excitation and the systems response thereto. In fact,
when j = jti) the transformed system equations describe the system's
response to a steady-state sinusoidal excitation or input signal. The real
part of the complex frequency, o (unit s"1), is associated with
transientsfor example, the rate at which the response to a step change
settles down to a steady state.
The real power of the Laplace transform is immediately apparent from
the transform of the rate of change or derivative df/dt of the function f{t):

= sF{s) - /tO). (14.2)

In the frequency-domain, multiplying the transformed function F(s) by


s is equivalent to differentiating the original function J[t) in the time-
domain. The initial condition J[0) must be added to complete the
operation, but 5 can be thought of as an operator in the .9-plane or
frequency-domain, that represents differentiation in the time-domain. In
a similar fashion, the operator ]/.t represents integration in the time-
domain.

14-2
14. C o n t r o l systems perfo rm a n ce

So m e c o m m o n tran sform s are given in Table 14.1.

JW *U)
unit impulse S (<) I l
unit step 2 z
s z- 1
1 z
O '" s*a z - e~ T
cos co0/ s 2{z~ cos 0l> T)
S2 * 0)02 z 2 -2 z cos (J T* 1
<>0 z sin u T
sin ai0t z 2 - 2z cosu>T+ 1
4-Z + 0>02
s+a X / e 'sTcos(i) 7)
e~* cos Ci)</ (s + a)2 + (002 z 2 - 2ze~aTcos a) T+ e'2aT
e 'at sin G)0/ o ze~Brsm<i>r
(s + a)2 + co02 z 2 - 2 ze'aTcosu> T + e 2fl3"
Table 14.1 T able o f L aplace a n d z tra n sfo r m s

The time response f(t) can be recovered from the frequency-response


function F(s) by applying the Inverse Laplace transform:

A t) = ~2irj [ T f i s ^ d s (14.3)
In many cases this contour integral does not need to be evaluated: instead,
the function F{s) that is to be inverse-transformed is expanded by partial
fractions into a collection of simpler expressions that are in a standard
table like Table 14.1. After inverse transformation of the partial fractions,
the resulting time functions are added together to get the overall time
function required.

14-3
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

*<8

Fig. 14,1 Step response associated with various pole positions in the complex plane
14-4
14. Control systems performance
14.2.2 Transfer functions
A transfer function represents the relationship between the driving
(input) signal and the response (output) of a dynamic system. It is
defined as the ratio of the Laplace transforms of the system output and
input, assuming zero initial conditions. With this definition, the transfer
function concentrates on the relationship between one output and one
input. If there are many inputs and many outputs, there will be a
different transfer function for every pair of input and output signals.
The roots of the denominator of the transfer function are known as the
system poles. (See also section 14.5.5). They are often plotted on the
complex plane as in Fig. 14.1. They describe the dynamic characteristics
of the system. If any pole has a positive real part a, it means that there
is an exponentially increasing element in the response and the system is
unstable. If the poles have non-zero imaginary parts, it means that the
response is partly oscillatory. This can be seen in Fig. 14.1. Negative real
poles indicate an exponentially decaying response. Complex poles always
come in pairs, with equal and opposite imaginary parts, i.e. conjugates.
14.2.3 Example of a DC or brushless DC motor
In Chapter 1 it was shown that the DC and brushless DC squarewave
motors can be represented by the same simple model in which V
represents supply voltage, /is armature current, / is armature resistance,
kE is the EMF constant, kj is the torque constant, Tc is the motor torque,
J/ is the rotor inertia, a is the rotational acceleration, and gjm _ is the
angular velocity. Armature inductance is assumed to be zero and any
mechanical shaft resonance is ignored.
We have seen in Chapter 1 and elsewhere that the speed is essentially
controlled by the voltage: in the steady state, assuming the load torque
is zero, the speed is given by equation (1.7) as

This steady-state equation gives no indication as to how the speed varies


dynamically if the supply voltage V is changed, as it might be by the
action of the electronic controller. The dynamic response can be
determined from the transfer function between speed and voltage, which
is derived as follows from the equations that describe the system.
14-5
D e s ic n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

V= RI+ *E m; (15.4)
Tt = k , /; (14.6)
re = /a - (14.7)
and
(14.8)
dt
- a .

Assuming that the speed is zero at time t = 0,


a = (14.9)
so that
to m h = h i
Js Js (14.10)
kj (V ~
*7?
This can be rearranged to give the transfer funcdon between speed and
voltage as
ttm(*) = IMk _ 1/*E
Kj) j + RJ 1 + sT (14.11)
S kjkrE
The denominator has only one root, so the transfer function has only
one pole at s = -l/T , where T = RJ/kjky, and the system is said to be a
first-order system. T is called the mechanical time constant After a
disturbance such as a step-change in voltage, the speed would be
expected to settle exponen dally to a final steady-state value. This is
because the transfer function is of the form 1 /(j + a), and from Table
14.1 this has an inverse transform that is a decaying exponential function
of time. After an infinite time, in the steady-state all derivatives are zero
and the corresponding condition in the frequency domain is s = 0. If this
is substituted in equation (14.11) the >term in the denominator
disappears and the transfer function degenerates to the steady-state
expression l/Ag, as in equation (14.4). This value is callcd the gain of the
system.

14-6
14. C o n t r o l system s p e r f o r m a n c e
14.3 Modelling drive components
In developing a controller for a servo system, it is necessary to create a
model of the "plant", which is that part of the system consisting of the
motor, the drive, the position and/or velocity transducers, the load, and
any mechanisms such as actuators or gear trains. Most control system
design methods require a linear plant model. This means that the
coefficients in the differential equations are constant: i.e., independent
of time and of speed, voltage, etc. If the plant model is not linear, it can
sometimes be linearised for small variations about an operating point
using "perturbation analysis.
14.3,1 Brushless PM motor model including inductance
The simple dynamic model of the brushless DC motor can be extended
to include the effects of inductance as follows. The electrical equivalent
circuit is shown in Fig. 14.2. E is the back-EMF, equal to The
electrical equation of the motor is
V =L ~ +R I + (14.12)
dt b m
We can assume that the magnetic field is constant and that the
electromagnetic torque T is proportional to current:
T =V - <14-13)
The mechanical properties of the motor are the inertia Jm and friction
torque Tf. Friction is often a nonlinear function of speed, and it is usual
to allow for a "viscous damping term" Dv>m, to represent at least that part
of the friction that is proportional to speed. The load can often be
Ia R La

fig- 14.2 Equivalent circuit of a brujhlcss DC motor

14-7
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a c n e t m o t o r s

described by a constant opposing torque TL and an additional inertia


The mechanical equation of the motor is
T = (/m + / L ) ~ - * D u * Tt + Tl . (14.14)
at
To obtain the transfer function between speed and applied voltage, the
differential equations (14.12) and (14.14) are transformed into the
frequency domain by the Laplace transform. From equation (14.12),
V= (L s +/?)/ + kzu m. (14.15)

Substituting equation (14.13),

T = <V ~ (14.16)
Ls + R
Neither friction torque, 7} nor load torque TL affects the motor transfer
function, so they can be set to zero. The mechanical differential equation
(14.14) then transforms into
T = (/m * JL) s u m + (14.17)

Equations (14.15) and (14.17) then give


V - kEk r com = (L s + R ) m [(/m JL)s * D]. (14.18)
From this, the transfer function between speed and voltage is

(14.19)
\Xs) (L s * * )[(/ + f j s + D] * kEk ,

14.3.2 Mechanical and electrical time constants


The poles of the system are the roots of the denominator of equation
(14.19), which is quadratic. There are therefore two roots, and it is
common to write equation (14.19) in a form which separates these two
roots: for example,
14-8
14. Control systems performance

6*.y) = = -----------]lh ---------- (14.20)


K-ff) ( s r . * l ) ( f T _ + 1)

where t c is the electrical time constant and Tm is the mechanical time


constant As before l/k E is the system gain. Assuming no motor damping
and low inductances, the time constants are

Equation (14.14) assumes that the shaft coupling Jm and JL is torsionally


rigid: in other words, its torsional stiffness is infinite and there can be no
torsional oscillations between the motor and the load. Any real shaft
system will flex under the torsional stress induced by load torque and
acceleration torque. To model the extra dynamics associated with
torsional deflections of the shaft, additional differential equations must
be added, representing the rotors and the shafts that join them by a
series of inertias and spring constants. The load and motor speeds and
angular positions will not necessarily be the same, and there may be an
additional phase lag between motor and load rotation that will affect
performance especially when rapid speed changes and reversals are
required. This is important not only in the performance of the servo
system, but also in the design and application of the shafts and couplings
themselves, because oscillatory torsional vibrations tend to have low
damping and a magnification of torque can occur if resonant torsional
frequencies are excited. This can lead to failure of shafts or couplings,
either in a single catastrophic event or by fatigue. A simplified and very
clear analysis of torsional resonance is given in Ref. [1].
Other unmodelled dynamics include nonlinearities in the drive circuit,
magnetic nonlinearities in the motor, and variations of load torque as a
function of times, speed, and/or position. In some cases these effects arc
too difficult to treat by simple transfer-function analysis, and full-scale
simulation is necessary. For this purpose there exist many computer
programs such as SABER, SIMUUNK, MATr ix -X , EASY-nVE, EMTP,
CONTROL-C, ACSL, and others.

14-9
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

14.3.3 Transducers
When a motor is used as part of a control system, it is desirable to
measure some output variables of the system for comparison with the
desired values. Velocity and position are common examples. In position
servos, it is usual to have feedback of both position and velocity. One
solution is to measure position using an encoder and to use electronics
to calculate velocity from the encodcr output. Alternatively, resolvers may
be used with resolver-to-digital converters.
14.3.4 Load effects
The load affects the performance of the control system. For example, a
high load inertia JL slows the response, and a variable load torque can
cause unwanted speed variations. It is therefore important to account for
the load when modelling the system and to design the controller to be
sufficiently robust to allow for variations in the load.

14.4 Control systems


14.4.1 Feedback and closed-loop control
The transfer functions in equations (14.11) or (14.19) are upen-loop
transfer functions, because they do not incorporate any feedback of the
speed for comparison with the desired value.
With opcn-loop control, a step change in voltage Vwill cause the speed
to change until it reaches a new steady-state value. A serious disadvantage
of open-loop control is that the speed can vary as a result of disturbances
such as changes in the load torque. With open-loop control and no
feedback, the controller has no information about such speed variations
and therefore it cannot take compensating action.
An equally serious drawback of open-loop control is that the output
variable (speed) depends on the gain as well as the timc-constants of the
system. Both of these can change as a result of temperature-variations,
ageing, etc.
In an ideal control system, the controlled variable (speed) would be a
function of the demand signal (voltage), and independent of everything eke.
14-10
14. Control systems performance
T h e degree of independence from extraneous influences is known as
robustness. To ensure robustness, all practical control systems involve
feedback.
Feedback means comparing the output variable (speed) with the demand
signal that represents the desired speed. In an analog system the actual
speed and the demand speed are both represented by voltage signals and
the comparison is effected by subtraction using an op-amp circuit. In a
digital control system the actual speed and the demand speed are both
represented by digital words, and the comparison is effected in a micro
controller. In both cases the difference between the desired speed and
the actual speed is used to "drive" the plant. The difference signal is
called the error, and the feedback mechanism ensures that the error is
minimised: in other words, it forces the output variable to approach the
desired value. The higher the open-loop gain, the smaller the error and
the closer the output variable approaches the desired value. The element
in the controller that performs the subtraction or comparison is referred
to as a summing function or error amplifier.
The signal representing the desired value of the controlled (output)
variable is called the reference, for example, the speed reference. When the
comparison of the actual signal with the reference signal is a subtraction,
the feedback is called negativefeedback. The signs of the signals connected
to the error amplifier are important. If one of them has the wrong
polarity, the feedback becomes positive and instead of driving the output
to minimize the error, the system becomes unstable.
A simple feedback system is shown in Fig. 14.3. W(j ) is the open-loop
ERROR SIGNAL

14-11
D esign of brushless permanent-magnet motors
transfer function of the plant, and o rcf is the speed reference signal at
the input. The feedback loop tends to reduce the error between the
actual motor speed and the reference speed. Often the error signal is
amplified to increase the gain, as shown in Fig. 14.3. The amplification
may include "dynamics", such as the parallel addition of a term
proportional to the integral of the error signal. Accordingly the error
amplifier has its own transfer function C(j) as shown in Fig. 14.3. The
controller C(s) is designed to optimise certain system properties such as
robustness, stability and bandwidth, and it is sometimes known as a
compensator.
The transfer function of the closed loop system shown in Fig. 14.3 can
be derived easily as follows. If the error signal is represented as E(s) then
from Fig. 14.3
r(s) = 0 W{s) x B{s) and E{s) = U(s) - Y(s) ( 14-22)
so that by eliminating (j),
n s ) _ a s) m s) (1428)
U(s) 1 + a s ) W {S)'
This expression is written in terms of a general output variable Y(s) and
a general reference signal U(s). The denominator is significantly changed
by closing (he feedback loop, and so also is the gain. The system poles,
and therefore all its response characteristics, can be cxpected to be
significantly modified by loop closure, and later sections will examine
how these changes can be controlled.
14.4.2 Speed controls and servo systems
A speed controller that controls forward speed only, and is not capable
of producing a negative or braking torque, is known as single-quadrant
control, Fig. 14.4. It operates in only one quadrant of the spced/torque
plane. A true servo system operates in all four quadrants. That is, it can
produce motoring or braking torque when the motor is running in
either direction.
A control system alters a dynamic system to give more accurate control
of speed, position, etc., and it is often required to meet other
performance objectives such as robustness, stability, and rapid response.
14-12
14. Control systems performance

Fig. 14.4 Definition of single- and four- quadrant motor control


A feedback control system works by forcing the output lo follow the
commanded input closely.

14.4.3 Speed control


A speed signal can be obtained direcdy from a tachometer or resolver,
or it can be synthesized from a train of encoder pulses. It can then be
compared with a commanded speed reference signal. Speed control is
the commonest form of motor control.

14.4.4 Position control


Position is measured by an encoder or potentiometer and fed back for
comparison with a position command signal. Many position controllers
also include a velocity loop which tends to add damping, thereby
improving the stability of the system. The resolution of the position
control system depends on the sensor. 1000-line encoders are popular in
many applications. With two quadrature channels there are effectively
4000 signal edges per revolution, giving a resolution of 360/4000 = 0.09
degrees. This is not the same as the accuracy of the encoder, which
depends on the precision of the encoder disk and optical system, and not
merely on the number of pulses per revolution (PPR). Similarly with
resolvers, there is theoretically no limit to the resolution because the
signal from the transducer is an analog signal; but the accuracy of this
signal is less than 100% because of imperfections in the resolver itself
and in the decoding circuits. It is quite possible for a resolver system with
12-bit resolution (i.e. 212 = 4906 counts per revolution) to be less

14-13
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

accurate than a 1000-line encoder, even though the resolution is about


the same in both cases.
In position control systems, the position transducer may not be on the
motor shaft, but remotely located, particularly if there is an actuator, gear
train, or some other mechanical linkage between the motor and the
controlled object. Mechanical linkages of this type introduce backlash,
hysteresis, and resonances into the system. All of these make the analysis
more complicated, but the principle remains the same, that the feedback
loop "surrounding" the plant (which now includes the mechanical
linkage) tends to make the system response independent of the
imperfections of the plant.
14.4.5 Torque control
Assuming torque is proportional to current, torque control is effectively
the same as current control. Torque control is important in applications
like electric vehicles, where pedal pressure represents a torque demand
signal and the function of the entire speed feedback loop is performed
by the driver, who compares an internal "desired speed" with the actual
speed indicated by the speedometer (or even estimated visually). The
torque can usually be changed rapidly because the current can be
changed rapidly. This is in contrast with speed control, because the rate
of change of speed is slowed by the mechanical inertia.
In some applications, constant speed must be maintained even though
the load torque varies widely and suddenly, and in these cases the ability
to change torque quickly is very im portant An example is in machine
tools: when the cutting tool first makes contact with the workpiece, the
load torque at the motor shaft increases suddenly, but the speed must be
maintained at a steady value, otherwise the machined surface may
acquire a poor or inaccurate finish as a result of variations in cutting
speed.
14.4.6 Incremental motion control systems
It may be that more than one control mode is necessary. This can
happen when, for example, the motor must move between set positions
at a constant velocity. Such a controller can be realised by putting
switches in the control loop together with a switching circuit for
actuation. These switches are generally implemented in software. A block
14-14
14. C o n t r o l system s p e r f o r m a n c e

ERROR SIGNAL

Fig. 14.5 Block diagram of a hybrid position and velocity control system

diagram for a switchable velocity/position control system is shown in Fig.


14.5. This is only a simple example. Modern incremental motion control
systems have sophisticated capabilities including the ability to follow motion
profiles and to communicate with high-level computer controllers and
PLC's that may be controlling tens, hundreds, or even thousands of
motor drives in one installation.

14.5 Characteristics of closed-loop control systems


The characteristics of a closed-loop control system include the step
response, frequency response, bandwidth, steady-state and stability
margins, etc., and these can all be used to specify the desired response
of a controlled system.
14,5.1 Frequency response
The frequency response describes the response of a system to a sinusoidal
input signal as the frequency of that signal varies from zero to infinity.
14-15
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

If the system is linear, a sinusoidal input stimulates a sinusoidal output


of the same frequency. As the frequency of the input signal is swept
across the spectrum, the magnitude and phase of the output signal vary.
The frequency response is often illustrated using a Bode diagram, which
is a pair of graphs graph showing the variation of the gain and phase vs.
frequency. The gain is generally defined as the ratio between the output
signal and the input signal, and the phase is the phase difference
between them.
The Bode diagram can be calculated by substituting the expression j-j(o
into the transfer function, separating the magnitude (gain) and phase,
and plotting the resulting functions vs. frequency 0). On a Bode diagram,
frequency appears on the x-axis in Hz or rad/s, usually with a logarithmic
scale. Magnitude or gain is plotted on the y-axis, cither in ordinary units
or in decibels (dB ).1 Phase change is plotted in degrees.
The Bode diagram for the first-order motor transfer function of equation
(14.11) can be seen in Fig. 14.6. As the frequency increases, the gain
decreases. When the frequency &> reaches k^k^/R JnA /i, the gain is 1/V2
of the DC or zero-frequency value, i.e. and its decibel value is -3 dB. Fig.
14.6 shows that the phase of the output signal lags behind that of the
input signal. The phase shift is exactly 45 at the pole frequency. Note
that the phase shift is perceptible at frequencies much lower than the
pole frequency. The phase shift is often a more sensitive indicator of the
ability of the system to follow modulating signals.
Bode diagrams are used for stability analysis and for control system
design. They are particularly useful because they can be generated
experimentally for a completely unknown system, by forcing that system
with sinusoidal input signals at different frequencies and plotting the
gain and phase of the transfer function between the output signal and
the input signal. The resulting plot can be used to identify a model for
the system. System identification means extracting values for the poles and
zeros from the plotted frequency-response data. The experimentally
determined poles and zeros can be used subsequendy for stability analysis
or in control system design.

* The dccibel i> defined as 20 !og]n((7), where G is the ordinary value of the gain. Some
normalization may be necessary to express the gain in dimensionless units or per-unit
before calculating the decibels.

14-16
14. CONTROL SYSTEMS PERFORMANCE

PHASE U G H

Fig. 14.6 Bode diagram of a first-order motor drive system with a single poic at
-*i.kt/RJ.

In the motor drive system the Bode diagram could be measured by


modulating the applied voltage ^sinusoidally, in other words, by adding
a small sinusoidal variation of constant magnitude and variable frequency
to the steady DC value. The speed would then vary sinusoidally. The
tachometer signal representing speed is a voltage that can be compared
with the signal-generator voltage that is exciting the sinusoidal
component of the gain would most conveniendy be defined as the
ratio of these two signals, normalized to the DC value of the same ratio.

14-17
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a c n e t m o t o r s

14.5.2 Bandwidth
The bandwidth of a system is the frequency range over which the gain is
within 3 dB of the maximum gain. In servo drives the maximum gain is
usually the DC gain, reflecting the fact that slow variations in the demand
for velocity or position changes can be followed with high accuracy and
little phase shift. In order to follow rapid changes in the demand signal1
with high accuracy and minimal phase shift, the system requires a high
bandwidth. The bandwidth is generally different for different controller.
A torque control can generally have a much wider bandwidth than a
velocity controller, mainly because of the effect of inertia in slowing
down the rate at which speed changes can be made. A position controller
generally has a lower bandwidth. A very high velocity-loop bandwidth
requires low motor inertia, and the ultimate performance is achieved
with "moving coil" motors or ironless motors, which can have a
bandwidth of several thousand rad/s. Position loop bandwidth is
generally below 1000 rad/s.
A high bandwidth may imply a high power requirement if high torque
is to be repeatedly applied at high velocity. This directly affects the cost
of a servo drive. As the design bandwidth increases, so does the system's
susceptibility to noise and interference, requiring special measures in the
screening of transducer signals and the electronic circuit design.
14.5.3 Step response
It is often of interest to see the time-domain response of a dynamic
system to a step change in input signal. A step input contains information
at all frequencies, is often easy to apply experimentally, and gives an easy-
to-understand illustration of the how the system behaves. The response
can be calculated using the inverse Laplace transform (section 14.2.1).
For a first-order system such as the motor described in equation (14.11),
a step input is represented in the s-domain by the function V/s, and the
response is the speed as a function of time, given by
V Itffc
s (14.24)
1 + s--&RJ. J

14-18
14. CONTROL SYSTEMS PERFORMANCE
U sing partial fractions,
n ( ') = (14.25)
s + ^T-^E
RJ
From Table 14.1,
m<') - IT fi - '* ] (14.26)
w here

is the mechanical time constant. This time response can be seen in Fig.
14.7. It shows immediately how long it takes for the system to setde into
a new steady state. After one time constant (t = t ), the function 1 - e_1
* 0.632, so the speed is within 36.8% of its final value. Table 14.2 shows
the approach of the speed to its final value in these terms. After 5 time
constants, the speed is within 0.7% of its final value. This setding time
might be too long in many applications, and feedback controllers may be
designed to speed up the response.

NORMALISED OUTPUT

14.7 Step response of motor drive with first-order transfer function

14-19
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a c n e t m o t o r s

/A 1 - e U'
0 0 1
l 0.632 0.368
2 0.865 0.135
3 0.950 0.050
5 0.993 0.007

T a b le 14.2 E x p o n e n t ia l a p p r o a c h t o a steady -s t a t e value


FOLLOWINC A STEP-CHANCE INPUT

14.5.4 Stability; gain and phase margins


A system is said to be stable if the output converges towards a steady state
following a disturbance or change in demand signal. For linear systems,
stability can be determined from the transfer function, but for nonlinear
systems the stability can also depend on the magnitude of the input
Sometimes a nonlinear system may be stable for small disturbances but
unstable for large ones. In an unstable system the output increases or
diverges, often in an exponential or oscillatory manner, with loss of
control and potentially catastrophic results.
Obviously stability is a very important criterion for any closed loop
control system. To ensure that the closed loop system will be stable even
for variations in plant parameters and external disturbances, it is usual
to specify measures of relative stability. Two commonly used measures are
gain margin and phase margin. The gain margin is the amount of
additional gain required to make the system just unstable. The phase
margin is defined as the amount of additional phase lag required to
make the system just unstable. The margins can be determined from the
system frequency response graphically or from the system transfer
function. They can be used as part of the closed-loop system specification
to ensure that the system remains stable. They are based on the fact that
if the gain has a magnitude of 1 and a phase shift of 180, the system will
be just unstable if the output is fed back and subtracted from the input.
In a negative-feedback system this subtraction would produce an error
signal of 2 per-unit. Because of the 180 phase shift the error signal can
never be reduced to zero and therefore a steady state in which thf
14-20
14. Control systems performance
output equals the input or demand signal can never be reached. The
gain margin can be read from the Bode diagram as the amount of
additional gain required to reach the ideal value of 0 dB at the frequency
where the phase-shift is 180. The phase margin is the amount of extra
phasc-Iag required to reach 180 at the frequency where the gain is 0 dB.
14,5.5 Steady-state error
The final-value theorem states that
lim f{() = lim [s^-s)]. (14.28)
w s -* 0
It can be used to calculate the steady-state value of a system whose
transfer function is F{s). Referring to Fig. 14.3, suppose C(s) = 1 for the
sake of simplicity, and let W(.f) = G/( 1 + rrm). This is the open-loop
transfer function, and it is first-order, with a single time constant x . The
transfer function between the output K(j) and the input U{s) is given by
eq u atio n (14.23) as

1 + H is )
= 7 :--------------U.S)
(1 + G)
+ jr m
(14.29)
while the transfer function between the error signal and the input is

B(s) = ----- i----- U(s) = ------1 + ----. (14.30)


1 + W {s) (1 + G) * s r m
Using the final-value theorem, the steady-stateresponse reached
an infinite time afterapplication of a step-changeof unit magnitude in
u(t), is given by the limit as s - 0 of [j l / s F(j)]. From equation
x x

(14.29) this is
09) = (14.31)
\*G
Similarly for the error e{t), the limit as s - 0 of [ j x l / s x (.t)] is
* ' - > ) = 1 + Cr (H.32)

14-21
D e s ic n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

This result reveals an error in the final value of the system with respect
to the reference input. The steady-state error decreases as the forward gain
G of the system is increased, but to remove the steady-state error
completely, it would require infinite gain which is not practical. In
practice, increasing the gain generally produces instability sooner than
it reduces the steady-state error to an acceptable value.
14.5.6 Integral gain compensation
If the error amplifier has a transfer function C(.t) = 1/s, it is an integrator
rather than a plain amplifier, and the open-loop transfer function
acquires a pole at the origin (s = 0):
a s) ^ 5 ) = - -----(14.33)
* 0 + "m )
The effect of this on the steady-state output signal is that

yU ) = lim = 1. (14.34)
o sU + sr m) + G
Similarly
e (t >) = slim * ^m ) (14.35)
o j( l + ^rm) +G
Equations (14.34) and (14.35) show that the integrator removes the
steady-state error completely, A system whose open-loop transfer function
has a single pole at the origin is called a Type 1 system, and it has a zero
steady-state error. With no pole at the origin, the system is a Type 0
system and there is a finite steady-state error that depends on the gain.
14.5.7 Root locus
The characteristic equation of a system is the denominator of the transfer
function equated to zero. The roots of this equation are the system poles
and they reveal much about the system dynamics. In a closed-loop
feedback control system such as that illustrated in Fig. 14.3, the
characteristic equation is obtained from equation (14.23) as
1 + CW = 0. ( 14. 36)

14-22
14. C o n t r o l system s p e r f o r m a n c e
M the controller parameters are varied, the values of the poles change.
T hese poles can be plotted on the complex plane as a function of some
design parameter of the controller (e.g. the gain G), and the resulting
g raph shows the migration of the poles as this parameter is varied. The
set of graphs for all the system poles is called the root locus diagram.
A typical root locus for a third-order system is shown in Fig. 14.8. For the
sake of argument, assume that the variable control parameter is the
forw ard open-loop gain term G. We can tell the number of roots
immediately from the number and nature of the branches in the
diagram . The leftmost branch is associated with a single real pole that
starts at point E when the gain is low. As the gain increases this pole
migrates leftwards along the negative real axis. The right-hand root locus
begins with two negative-real poles at points A. As the gain is increased,
these poles approach one another until they merge at C. At higher values
of gain they separate into two conjugate poles with an imaginary part that
indicates that the system has an oscillatory component in its response.
Ihe system is stable provided the gain is not too large, as long as the
poles retain a negative real part. At higher values of gain the complex
poles migrate into the real half-plane and the system is then unstable. A

14-23
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

step input would excite an oscillatory response with exponentially


increasing oscillations. (See Fig. 14.1). A well-behaved system typically has
all its poles in the left half-plane and within 60 of the negative real
axis.
The root locus plot is a powerful design tool. In a detailed root-locus
diagram, values of the parameter G would be marked along the root loci
to show the designer what ranges of values are acceptable. The designer
can select a particular damping ratio (see next section) directly from the
plot and calculate the gain from that plot.
14.5.8 Second-order systems: critical damping
The roots nearest the origin are the dominant poles and these poles
determine the essential characteristics of the response. Two particular
aspects of the system response which are usually very obvious, particularly
in the step response, are the frequency of any oscillatory component and
the rate at which this component decays or is damped out. The lowest-
order system that has an oscillatory component in its step-response is a
second-order system such as the one described by equation (14.20). The
denominator of this transfer function can be written as
TeTm(y2 + 2 t u as+ ton2) (14.37)
where
co = and f = .M T e * (1
v're'rm ^
The system poles are the roots of the bracketed term in equation (14.37),

s = [-C j\/l - C2] wn- (14.39)


The parameter is called the damping ratio and Ci>n is the undamped
natural frequency of the system. If C < 1 the response is oscillatory, and if
>1, it is n o t (See Fig. 14.1). For ( > 1 the two roots are both real. The
case C = 1 is called critical damping. A critically damped system has no
overshoot, whereas an underdamped system ( < 1 ) has an overshoot and
at least one cycle of oscillation before settling. If the steady-state value of
the response is unity, the maximum value of the response during the first
overshoot is

14-24
14. Control systems performance
ym a - 1 . . - t h / T ? < 1 4 4 0 >

and this occurs at time t = n/a>nV(l - 2). The setding time is


approximately estimated as 4/{(an seconds, or four "effective time
constants".
For example, if the damping ratio is 1/V2 = 0.707, = 1.043. This
value of C causesthe poles to lie at 45 from the negative real axis. At
60 from the negative real axis the damping ratio is t = 0.5 and the
maximum value is ;y[[iax = 1.163, i.e., a maximum overshoot of 16.3%. In
the interests of rapid response, some degree of overshoot may sometimes
be tolerated and this helps to explain why control systems are often
designed with the objective of locating the poles somewhere between 45
and 60 from the negative real axis.

14.6 Control systems - design


Control systems are designed to meet specific performance criteria.
General criteria are
1. Operation with as litde error as possible.
2. A well damped non-oscillatory response.
3. Robustness to system parameter changes as well as external
disturbances.
The design criteria are often conflicting. For example, increasing forward
gain will decrease steady-state error but will tend to make the step-
response more oscillatory.
There are many approaches to control system design. Some of the
simplest methods involve lead/lag compensation. Compensators (Fig. 14.3)
can be designed using open-loop frequency-response data and can be
implemented using simple R-C circuits. The compensator adds phase
lead or lag. Root-locus design is a more direct approach to lead/lag
compensation. The response can be determined from the dominant pole
positions on the root locus plot. The root locus is reshaped by the
compensators to give the desired system poles. A more direct approach
is to specify the closed-loop transfer function directly. A transfer function
describes the complete closed-loop performance including steady-state

14-25
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

and transient behaviour. The control system alters the plant dynamics to
match the specified closed-loop transfer function. The specified transfer
function must be robust to disturbances and plant parameter variations,
and of course the controller must be realisable and economical to
produce.
All of the above methods require a certain amount of judgement in their
execution. There is usually more than one possible solution and some
compromises may be required, even including a certain degree of
arbitrariness. As control design techniques become more sophisticated,
however, this arbitrariness decreases. More complicated techniques are
often based around optimising some quantitative performance index or
'cost function'. This cost function could be the mean squared error
between system output and a desired response for example. It could also
be subject to constraints on steady-state error, rise-time or overshoot.
Such control design methods produce high-performance, robust
controllers which often require accurately measured system variables for
feedback and high bandwidth controllers. Both facilities are expensive.
14,6.1 Lead-lag compensation
Lead and lag compensators are among the simplest of controller designs.
They work by altering the phase of the plant. A transfer function of such
a compensator is given by
His) = 1 + sl* . (14.41)
1 + s/aa
For a>l the equation represents a lead compensator, and for a<l it
represents a lag compensator.
The Bode diagrams of a lead and a lag compensator are shown in Figs.
14.9 and 14.10. Lead compensators increase the bandwidth by increasing
gain at higher frequency, and they increase the stability by increasing the
gain and phase margins. Lag compensators can be used to decrease the
bandwidth or alter the phase and gain margin to a specification. They
can generally be implemented using R-C circuits. Design of lead and lag
compensators normally involves specifying closed-loop parameters such
as gain and phase margin and then using root locus, Bode diagram or
Nichols chart techniques (among others) to design a controller within
practical constraints such as not having any poles too near the origin and
ensuring that o in equation (14.41) is not too large.
14-26
14. CONTROL SYSTEMS PERFORMANCE
GAIN (dB)

LOG FREQ UENC Y


PHASE LAG O

LOG F REQ UENC Y


Fig. 14.9 Bode diagram of a typical lead compensator
GAIN(dB)

PH A SE LAG O
LOGFREQUENCY

LOGFREQUENCY
Fig. 14.10 Bode diagram of a typical lag compensator

14-27
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

14.6.2 Pole placement


If the system transfer function is known, the closed-loop system can be
specified by its transient response or a transfer function, and the
controller designed so that the closed-loop system has this transfer
function. This is illustrated by means of an example in which a DC or
brushless DC motor is to be used in a position control system with two
feedback compensatorsa gain in the velocity feedback path, and a gain
in the position feedback path, Fig. 14.11.
The first-order motor transfer function is
<*)( _ 100 (14.42)
i(s) 1 + 0.02 s
The transfer function of the motor with the two feedback loops as shown
in Fig. 14.11 can be calculated using equation (14.23). The resulting
transfer function is a that of a second-order system:
_ M = -------------------- 100-------------------- (14.43)
W.s) 0.02 j 2 + (100(7V + l ) j + 100 Gp
The position control system is to be designed according to a given
specification. This specification is a time to first peak of 0.04 seconds and
a setding time (within 2 % of final value) of 0.06 seconds for a step
response, thus allowing for a small amount of overshoot.

Fig. 14.11 Block diagram showing DC or brujhlcss DC motor with velocity and
position feedback

14-28
14. C o n t r o l system s pe r f o r m a n c e
As in section 14.5.8, the transfer function can be written in the generic
form for a second-order system:
iKs) = -^ (14.44)
.r2 + 2 C<on^ + o>n2
The normalized step response can be calculated using the inverse
Laplace transform (equation (14.3)) as

}{() = 1 ---- 1 e 'C'sin (conV/l - Z2 t + cos'*C). (14.45)

To find the timeto first peak, equate dy{t)/dt to zero: the result is
^ = -----------T = - (14-46)

Forthe setding-time criterion, for a 2% tolerance, the setding time is


approximately,
4~ ~ir~ (14-47)
C^n
Taking equations (14.46) and (14.47), the closed-loop system parameters
are
w - 103 rad/s

These values are then substituted into equation (14.44) which is equated
with equation (14.43) to give values for the control gains:
velocity gain Gv - 0.017 A/iad-s'1
position gain G - 1.89 A/rad

The transfer function for the complete closed-loop position control


system is:

14-29
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

-------- (o, = 103, ( = 0.65). (14.48)


As) s2 + 133 j + 10600

The poles are located a t.? = -67 j 78.3 and the closed-loop system step
response looks like Fig. 14.12.

0.00 0.02 0.04 0.06 0.08


TIME (SECONDS)
Fig. 14.12 Step response of closed-loop position control system showing response
within specified response limits

14.6.3 Robustness
Lead-lag compensation and pole placement are examples of good
classical control design techniques. One drawback is that they are
strongly based on the dynamics of the system being controlled and not
very robust to changes in the plant or to external disturbances. Dynamic
system parameters can change due to operating conditions and also due
to component ageing over the lifetime of the system. The load can also
vary, and sometimes one controller must be able to handle a variety of
different loads. The ability of a controller to cope with changes in plant
and load as well as with external disturbances is called robustness. More
robust controllers can cope with changes in plant dynamics and can
therefore be thought of as more general-purpose controllers. A well-
known and widely used robust control design is PID
con tTo\proportional/integral/denvatwe control.

14-30
14. C o n t r o l system s p e r f o r m a n c e

14-7 PID controllers


It is often the case that not much is known about the plant and load; or
that one controller, for economic reasons, is expected to be able to
control a range of systems. One solution to these problems is a three-term
or PID controller, so called because it consists of a proportional gain, an
integral gain and a derivative gain. The proportional term controls the
loop gain of the system and therefore reduces system sensitivity to
parameter variations. The integral term increases the order of the system
in order to reduce the steady-state error. The derivative term helps to
stabilise the system. The selection of the coefficients or gains associated
with these three terms is called tuning. It is often done by using an
established design method to get a starting point, followed by an intuitive
trial-and-error method. Control engineers use their experience in PID
design together with their knowledge of the plant to design a controller.
The method leads to an acceptable solution although not necessarily the
optimal one. It is often the best economic solution to the problem. More
advanced PID control can be implemented by adaptive control systems
which tune the controller "on line while it is actually controlling.
14.7.1 Design of a PID controller
The proportional term is a gain. Increasing this gain increases the speed
of response of the system and decreases the steady-state error, but it also
reduces the stability of the system and increases its susceptibility to noise.
If the plant is a Type 0 system, that is if it does not have a pole at the
origin, there will be a steady-state error in the output. The error can be
removed by adding an integral term to the controller, as shown in
section 14.5.5. The integration of the steady-state error tends to produce
an increasing value which the controller acts to eliminate by the normal
action of negative feedback. Only when the error signal is zero does this
action cease.
A problem with integral controllers arises from the fact that real
controllers are limited in the forcing functions they can apply to the
plant, and the plant itself has limitations: for example, the current
supplied to a brushless DC motor must be limited to a safe value and this
limits the maximum available torque. If the demand signal is too large,
integral error can ramp or wind-up to a value that calls for forcing
functions (currents, torques etc) that are larger than the safe limits. The
integral wind-up will have to be unwound at some point, possibly
14-31
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

interfering with the stability and speed of response of the system. To


counteract this, anti-windup can be implemented. At its simplest, anti-
windup limits the error-integral signal to a value that is sufficiently large
to remove the steady-state error in the response, but small enough to
avoid the problem of integral wind-up.
The derivative term gives the controller a predictive quality which tends
to stabilise the system. It guards against excessively large rates of change
and therefore increases stability. Derivative terms are difficult to
implement. The gain of a derivative term increases with frequency, so
high-frequency noise and interference (which is always present in real
systems) can be amplified to a potentially dangerous degree. Because of
this problem, especially with analog circuits, differentiation elements
should be avoided if possible.
In motor drive systems an alternative is to use velocity feedback in a
position control system, usually in a nested loop within the position loop,
although this can be expensive if another transducer is required. Another
solution is to incorporate a low-pass filter in the derivative term. This
ensures that the desired stabilising effect remains within the bandwidth
of the system being controlled, but high-frequency noise is filtered out
of the signal before it is differentiated.
14.7.2 Tuning a PID controller
There are many methods for tuning a PID controller, varying in
complexity from simple rules based on open-loop response to self-tuning
algorithms. One of the better known techniques is the Ziegler-Nichols
step response method.
In this method, a step response of the open-loop system is recorded. An
example is shown in Fig. 14.13. Two parameters are taken from this
responsethe maximum gradient, A, and the point at which this line of
maximum gradient crosses the time axis, B. The magnitude of the input
step is taken to be U. The three gains for a full PID controller are
calculated using
4 = 1.2-2- (14.49)
^ AB
7j = I B (14.50)

14-32
14. C o n t r o l system s p e r f o r m a n c e

Fig. 14.13 Step response of third-order system showing Ziegler-Nichols PID design
technique
7J = I B (14.50)
rd = 0.55 (14.51)
where the PID controller is represented by
a s) = ^(1 + 1!T{s + 7dj). (14-52)

The engineer would implement the controller with these gains and then
"adjust on test" to achieve a satisfactory closed-loop response.
14.7.3 Auto-tuning
It is possible to tune PID controllers automatically. An auto-tuner applies
certain inputs to the closed-loop system and records the output. From
this data, it calculates the parameters for the PIDcontroller. A potential
problem is that the input may exceed the plants limits of safe operation.
An alternative is adaptive control This means that the parameters are
continuously varied while the system is in operation.

14-33
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

14.8 Digital control


Most motor control is now implemented in digital electronics. Digital
controllers tend to be more accurate, less susceptible to noise, and more
flexible in terms of programming. The cost of digital control has fallen
dramatically because of the extraordinary development of integrated-
circuit technology. Digital controllers also permit a wide range of features
such as diagnostic information, self-calibration, communications, and
protective functions that would now be prohibitively expensive to
implement by traditional analog methods.
A digital PID controller can have features like anti-windup integral
action, derivative filtering, direct encoder input and summing junction
with analog output all on one chip. These devices are often programmed
with gains and other control parameters as if the controller is an analog
devicc. This means that classical continuous control system design
techniques such a5 those in section 14.6 can be used.
Modern microcontrollers and digital signal processors take the functions
to an even higher level of sophistication by providing powerful in-built
functions such as PWM output signals, sine/cosine lookup tables for
vector control, and even high-speed floating-point computing facilities.
These devices can sometimes be programmed using high-level languages
like C', which permits very complex control systems to be implemented
entirely in software. Special compilers are written for the most important
families of controller, to convert high-level instructions from C into
assembly-language. Moreover, the operation of these controllers can be
simulated in great detail long before the hardware design is finalised,
and in some cases simulation code written in 'C' can be downloaded to
the hardware itself to become the operating firmware in the controller.
14.8.1 Discrete system theory
The controller may be digital, but the plant is almost always analog. This
means that it is necessary to convert analog signals to digital and vice
versa. A common example is the conversion of a linear position
transducer signal into a digital word. This is done by analog-to-digital
converters (ADC) and digital-to-analog converters (DAG). Resolvers
produce an analog signal that is converted to a digital count by means
of a resolver-to-digital converter (RDC). ADCs and DACs are often
integrated into other chips such as motor controllers.
14-34
14- CONTROL SYSTEMS PERFORMANCE
Digital components work with a specific word length. The word length
is the number of bits used to represent each value. It determines the
resolution of the system and the smallest representable change in voltage.
For example, for a 12-bit ADC with an input range of -5 to +5 volts, the
smallest step is
5 ~(' 5) = * 0.0025 Volts. (14.53)
21Z 4096
The sampling rate is the frequency at which the analog signal is sampled
and converted to a digital signal. This rate is determined by the slowest
componentusually the controller chip. A higher sampling rate means
that the bandwidth of the controller can be increased and the system can
function at higher frequencies, but it may also mean a more expensive
controller. With a sampling frequency of N Hz the highest-frequency
signal that can be unambiguously sampled without aliasing is N/ 2 Hz,
known as the Nyquist frequency.
A sample of an analog signal effectively exists for an instant of time every
sampling interval. In a discrete system, there is nothing in between these
instants. When the signal is converted to an analog signal, the intermedi
ate values must be generated. The simplest method is the zero-order hold,
which holds the signal constant until the next sampling interval as in Fig.
14.14. A first-order hold uses the two most recent samples to generate a
slope between samples. This is more expensive. Higher-order holds are
rarely used in automatic control systems.
SIGNAL

TIME (SECONDS)
fig 14.14 Example analog signal with digital sampled signal and a zero- order hold
(ZOH)

14-35
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

14.8.2 Z transforms
Whereas the Laplace transform is used for continuous system analysis, the
z transform is used for discrete systems. It works by replacing a
continuous signal by a sequence of equally spaced impulses. The equal
spacing is determined by the system clock rate or sampling rate, which
applies throughout the system. The z transform is defined as
00
/=H = f[nT )z'a (14.54)
ZbO

where T is the time between successive samples. Some examples of z


transforms are given in Table 14.1. For systems analysis, it is necessary to
convert analogue elements, ADCs and DACs to the discrete domain.
The system can then be simplified using block diagram reduction
techniques to give a z transform input/output equation representation
of the system.
For the system in Fig 14.15, G(s) is the transfer function of the forward
path, i.e. the zero-order hold and the plant together:

G (s ) = I - e-sT \ (14.55)
s+a

ERROR SIGNAL, E(z) U(z)


DIGITAL DAC
* i |-ZnH
----- PLANT
R(z) CONTROLLER W(s) lY(z)

D(s)D(z) G(s)G(z)

Fig. 14.15 Block diagram of a typical digital feedback control Bystem

14-36
14. C o n t r o l system s p e r f o r m a n c e

Then
m = iA* a*)

=
Zf
1-^-11
\ s + a\
-
s (14.56)

= u m - z->)
Using partial fractions,

(14.57)
r(z) = (1 - z - l) U tf
ZHl-Zfcf
From Table 14.1,
Y(z) _ 1 - e -aT (14.58)
V{) z - e H i'
The controller equation is
U() = Uz ' 1 + AqE(z) (14.59)
C/jz) g V (14.60)
2?(2} z- 1

(4 = >w (14.61)

K(z) = D(z)G(z)[R(z)- Y(z)] (14.62)

n $ - W A G te (14.63)
R(z) 1+D(z)G(z)
Equation (14.63) is analogous to the continuous time expression for a
closed loop system in the s domain (equation (14.23)).
14-37
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

14.8 3 Z transforms and difference equations


The operator z can be thought of as the unit advance operator. This
means that the z-transform relationship
yit) = a y{z'x) + b y iz '1) + cu(z) (14.64)
is equivalent to
7R = + bA -i + c \ (14-65)
where
/ k - y(kD- (14-66)
Controller equations and digital filter equations are often given as
difference equations: then the number of memory locations required can
be seen at a glance. These equations can be transformed into z- domain
transfer functions which can be used for systems analysis or control
system design.
14.8.4 Stability of discrete systems
The z transform is related to the Laplace transform by the equation
z= (14.67)
This transformation maps the left half of the complex ^-plane on to the
unit circle in the complex z-plane. This means that stability and control
design methods developed for the continuous frequency and time
domains can be used for discrete systems with slight modifications. For
the root locus, the closed-loop system will be stable for poles inside the
unit circle. The damping ratio will increase as poles move towards the
origin. It is possible to map this circle back onto the left half plane with
the w transformation,
z +1 , z - w+1
w = ------- (14.68)
z- 1 w- 1

14.8.5 Digital control system design


A digital controller can be designed using continuous domain design
techniques by ignoring sampling and then converting the controller
equation into the discrete domain with z transforms as in section 14.8.2.
14-38
14. C o n t r o l system s p e r f o r m a n c e
This method assumes that the sampling frequency is high compared to
the highest natural frequency of the plant. A more direct approach is to
design it as a digital controller using a discrete control system design
method. An example is the deadbeat controller described in the next
section. A third option is to use a digital three-term or PID controller.
This is a popular solution and is described in section 14.8.7.
14.8.6 Deadbeat controller
A deadbeat or minimum-response controller responds very quickly with
zero steady-state error and a finite settling time. The step response of
such a system is shown in Fig. 14.16 and is a unit step delayed by one
sampling interval.
JW = z - 1 z - 1. (14.69)
This input is a step,

SPEED
ii
1.0- J t K * * X X X X

0.8 -
0. 6 -
0.4.
0 .2 .

0 I -------.----- -------.-------,-------,------- -------.-------*-------


0.00 0.02 0.04 0.06 0.08
TIME (SECONDS)

Fig. 14.16 Deadbeat step response of a discrete speed control system with sampling
time 7' 0.012 seconds

14-39
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

(14.70)
z- 1
so the discrete transfer function of the complete system is
na _ DiAGiA _ i {147l)
R(z) 1 + D(2>)G($ z

from equation (14.63). The controller transfer function is


DM 1 i z~l (14.72)
OM m G(z) 1 ~ z 1 .

Equation (14.72) is valid only for systems where the denominator is an


order of one higher than the numerator. The response given in equation
(14.69) is not realisable for other systems. A more general rule is to
design the controller based on a realisable output of
m = l (14.73)
- 1
where m is the difference between the order of the numerator and the
denominator of the z-domain transfer function.
The deadbeat response gets to the desired value exactly and does not
approach asymptotically as in a continuous-time controller design. It
should be noted that the response is accurate at the sampling instants
but not necessarily in between those instants. If the plant has high-
frequency modes, the signal could be oscillating between the sampling
instants.
This type of design is not very robust. If the model G(z) is not accurate,
the controller performance will deteriorate. In the case where the plant
is not known very well, a digital PID is a good robust solution.
14.8.7 Digital PID
For electric motor control, the PID controller is usually implemented
digitally. A typical digital PID algorithm is

1 4 -4 0
14. C o n tr o l systems pe r f o r m a n c e

u(") = ^ e(n) + i- Y , e(W + ^ e ( n ') - e(n - 1)]. (14.74)


AM)

Ap, kt and ftd are the proportional, integral and derivative gains
respectively. e(n) is the error signal or controller input at sampling
interval n and u(n) is the controller output.
The controller is protected against integral wind-up by having a limit
The integral sum is not allowed to exceed this value. Problems due to the
differentiation of noise are avoided by having a longer sampling interval
for the derivative term. The sampling interval means that the signals with
a frequency content higher than T J'i are ignored (where T is the
sampling interval). In this way, the derivative term has filtering action.
This type of digital PID controller typically comes in one integrated
circuit which can also include a summing junction and a digital-to-analog
converter for the controller output. It is normally programmed as if it
were an analog PID controller and the gains were analog PID gains.
14.8.8 PID control example
A nonlinear simulation program, PC-BDC [2] is used here to illustrate the
effect of a three-term PID controller on a 4-pole brushless DC motor with
surface-mounted permanent magnets. The motor is running at 2000 rpm
and is then forced with a step demand of a 200 rpm increase. After 0.04
seconds, the load torque increases from 0 to 0.45 Nm. Position is
measured by a 16-pulse encoder and the controller sampling rate is 2
kHz.
Fig. 14.17 shows the response of the system with no integral or derivative
termpurely proportional gain. The response to the step increase in load
torque shows a slight overshoot and a steady-state error of about 225 rpm
or about 10 % on the new set-point value of 2200 rpm.
Fig. 14.18 shows the effect of increasing the proportional gain. The
steady-state error is reduced but the overshoot is increased and the
response is slighdy more oscillatory.

14-41
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

lif t lo c it y (rp * ) u 1. Qo2

TiM (iic o n d f) k l.0*-t


Fig. 14,17 Response of a BDC motor with proportional control
U tlo e iu j <rp3 * 1 .0 * 2

Tint (f#cond> k l.0f-1

Fig. 14.18 Response of a BDC motor with a higher proportional gain

14-42
14. C o n t r o l system s p e r f o r m a n c e
Fig. 14.19 shows the system with a PI controller. The steady-state error is
reducing but the overshoots are large, leading to concern about the
stability of the system.
The output with a complete PID controller is shown in Fig. 14.20. The
derivative term stabilises the system by adding damping. The output of
each of the three stages in the PID controller is shown in Fig. 14.21 and
illustrates the effect of each of the components.
14.9 Advanced control techniques
PI and PID control are widely used in industry for motor control because
they are robust and give good performance at low cost. However, control
theory has advanced considerably beyond three-term controllers and
many new techniques have been developed which can be applied to
motor control. Most motor control system design is done using classical
techniques based on transfer function descriptions of the plant and
described elsewhere in this chapter. More advanced techniques are
available.
14.9.1 Adaptive control
Control systems are designed specifically for one plant and one
performance specification. Robustness can be designed in to allow for
variations in the plant, but only to a limited extenL Also, the model on
which the control system design is based must usually be linear.
Nonlinear models must be linearised at some operating point and this
also compromises the performance of the final system because the motor
is expected to operate beyond the range of validity of the linear model.
One solution is to design a controller which changes or adapts as the
operating conditions change, that is, adaptive control The controllers are
constandy changing and are often nonlinear. This makes it more difficult
to ensure that the system remains stable in all cases. Three common
examples of adaptive control are self-tuning regulators, model reference
adaptive control (MRAC), and sliding-mode controllers.
Self-tuning regulators are based on conventional controllers such as PI
but the controller parameters or gains are constandy varied according to
the performance of the closed-loop system.

14-43
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

U'laeity Crpmi Xi.OJ

rillt (fic o n d s) X 1. 0*-)


Fig. 14.19 Response of a BDC motor and a PI controller
lUlscilv <rp&> * 1.0*2

Tin* 'f 'C 0ds> k l.D t-1


Fig. 14.20 Retponsc of a BDC motor with a PID controller

14-44
14. C o n t r o l system s p e r f o r m a n c e
............. U3J...T. t ... I.Bem.__ ruTQC7.Bnc........... .

1.00
a.00 i
a.4Q 0.60 D.6D i.oa

...................EHTJEEBG&JREEIr..

cr n I rrzi ruYflw.m
.. li ftt-i.MSiLArf.fl.n
L.90
0*90
1.20
Capcorsdil *
1.00

fig. 14,21 Output of each of the three terms of a PID controller for a PID controlled
BDC motor

Model reference adaptive control involves feeding the input signals of the
actual controller into a dynamic model of the system that is simulated
continuously in real time on a computer. The input signals are
simultaneously fed to the real system, and the output signals from the
computer model are compared with the output signals from the real
system. The error or difference between these signals is used to adapt the
parameters of the controller continuously so that the closed-loop system
response is as close as possible to that of the specified dynamic model.
In sliding mode control, the state of the system (a vector defining the
velocity, position, etc) is constrained to move or slide along a predeter
mined trajectory in "state-space". This trajectory could be a curve on a
velocity/position plot, for example. The controller adapts as the point
representing the system state crosses and re-crosses the prescribed
trajectory, so that the system always approaches the prescribed trajectory
in a stable manner even if the system is actually unstable at points. It is
analogous to a bang-bang controller or to a hysteresis current-regulator.

14-45
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

It suffers from limit-cycles in the steady-state or chattering. Nevertheless,


it prodvices a very robust controller.
14.9.2 Optimal control
Many advanced control system design techniques work in the time
domain. They are based on optimising some "cost" function. This cost
function or figure of merit could be the integral of the absolute error
between output and input, or the integral of the squared error, for
example. Such integrals can also be weighted. By multiplying the absolute
error by time for example, as in equation (14.75), emphasis is put on an
accurate steady-state response:
F-$Jt\eW\dt. (14.75)
Other cost functions could incorporate maximisation of efficiency during
operation, or minimisation of acoustic noise.
The optimal control-system design problem is often formulated
mathematically as a constrained optimisation which is solved for a unique
solution, even for multi-input m ultioutput systems. This allows a control
system design which minimises the cost function while maintaining
stability and a degree of robustness.
14.9.3 Observers
A common problem in control is that the desired system outputs are not
available, often because the necessary transducers are too expensive. One
solution is to estimate the required state using a state observer. A state
observer uses a filter algorithm such as a Kalman filter to model the
system and statistically adjust the output according to what state variables
are available. An approximate value of the state variable is thus available
to be used in a feedback loop for a multi-input multi-output controller.
References
1. Electro-Craft Corp. [1980] DC motors, speed conlrolt, im systems
2. Miller TJE and McCilp M [199] ] High-speed PCrhased CAD far brushless motor drives, 4th
European Conference on Power Electronics and Applications, EPE 91, 435-439.

14-46
15. COOLING
15.1 Introduction
Heat transfer is as important as electromagnetic and mechanical design.
The analysis of heat transfer and fluid flow in motors is actually more
complex, more nonlinear, and more difficult than the electromagnetic
behaviour. It is often dealt with by means of simplified equivalent
circuits, and rarely receives the detailed analysis lavished on electromag
netic aspects.
Perhaps there is some justification for using approximate methods for
heat transfer, when exact methods are pursued for electromagnetic
design. The electromagnetic design determines the geometry of
laminations which are cut to fine tolerances. Their geometry and
thickness, together with their material properties and the design of the
windings, determine whether or not the motor will deliver the required
torque. They also determine the precise voltages and currents that will
be experienced by the power semiconductors in the controller. All of
these items critically affect the manufacturing cost. By contrast, as long
as the temperature rise does not exceed a nominal or specified value, the
actual thermal condition of the motor mainly influences how long the
motor will last, and has only a marginal influence on whether the torque
can be delivered. Furthermore, the motor designer often has little
control over the ultimate thermal environment of the motor, so there
may be little point in attempting exact thermal analysis.
There are two major aspects to the thermal problem: heat removal, and
temperature distribution within the motor.
In most motors heat is removed by a mixture of air convection,
conduction to the frame mountings, and radiation. In highly-rated
machines direct cooling by oil mist or even liquid coolants can be used
to achieve high power density. In "hermetic" motors used in refrigerator
compressors, the motor losses are usually transferred to the refrigerant
which may pass right through the motor.
The temperature distribution within the motor is essentially a diffusion
problem. It is difficult to analyze precisely, because of three-dimensional
' effects and "imponderable" parameters such as the thermal contact
resistance between, say, a bunch of copper conductors and a slot liner.
15-1
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a c n e t m o t o r s

Empirical rules are available, to be used with care. The most important
aspect of the temperature distribution problem is finding the hottest
temperature in the motor, given a certain distribution of losses and a
known rate of heat removal. The steady-state temperature distribution
can be very different from the transient distribution, and different
methods of analysis may be needed for the two cases.
The main reasons for limiting the temperature rise of the windings and
frame of a motor are:
1 to preserve the life of the insulation and bearings;
2 to prevent excessive heating of the surroundings; and
3 to prevent injury caused by touching hot surfaces.
The "life" of electrical insulation can be predicted only by statistical
methods, but in broad terms the life is inversely related to the
temperature, and the relationship is exponential, so that a sustained
10 C increase in temperature reduces the insulation life by approxi
mately 50%. Intermittent periods at higher-than-normal temperature can
be tolerated repeatedly, depending on their duration and the actual
temperatures reached. A spectacular example of this is the FUMEX
motor , which is used to extract fumes via the ventilation systems of
public buildings and concourses in the event of fire; these motors can
operate in an ambient temperature of 300 C for a limited period of 30
minutes. (After that the insulation life is well-and-truly used up). Similar
considerations apply to bearings. Grease-lubricated bearings maybe filled
with high-temperature grease for hot-running applications, but in
aerospace machines the bearings are usually lubricated by separately-
cooled oil or oil mist.
Heating of the surroundings is obviously undesirable especially if the
motor is heating the equipment it is driving. For this reason it is
important to minimize rotor losses, which are difficult to remove and are
conducted along the shaft. PM motors have cooler rotors than DC or
induction motors.

FUMEX is a trademark of Brook Crompton Parkinson Motors Ltd., England.


15-2
15. C o o l in g
In some applications such as hermetic compressors used in air-
conditioning, refrigeration, etc., the motor losses are removed by the
working fluid, reducing the thermodynamic efficiency of the system.
Exposed surfaces must be kept below 50 C (actual temperature) to avoid
injury or harm to people. In certain applications, for example under-
bonnet automobile auxiliaries, this requirement is impossible to meet
because the temperature under the bonnet may reach 100C. In
industrial applications the ambient temperature is generally less than
50 C, and NEMA ratings for electrical insulation assume an ambient
temperature of 40 C. In aerospace applications the ambient temperature
may not be as important as the temperature of coolant provided by the
airframe (oil or fuel); this may be as high as 200F.
The increase in winding temperature increases the resistivity of the
windings: a 50 C rise by 20%, and a 135 rise by 53%, increasing the f'R
losses by the same amount if the current remains the same. The increase
in resistance is used in test procedures to determine the actual
temperature rise of the winding, but this is obviously an average
temperature; hot-spot temperatures can be 10-20 higher.

15.2 Heat Removal


Heat is removed from a motor by conduction, radiation, and convection.
Usually the most important is convection of air. If the motor is flange-
mounted there may be appreciable conduction and consequent heating
of the motor mounting. Radiation is generally small but not negligible,
especially if the surface is enamelled, painted or lacquered black.
15.2.1 Conduction
The conduction equation for a block of thickness t and area A is
Q = kA^f m kA w (15.1)
ax t
where AT is the temperature difference through the thickness t. The
coefficient &is the thermal conductivity, with units (W /in2) per (C/in) or
W /C/in. [Sometimes this is written as W/C-in. The SI unit is W /C-
m]. The thermal conductivity is a material property, and usually it is a
15-3
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

function of temperature. Most metals have high thermal conductivities,


especially those which are also good electrical conductors. On the other
hand, electrical insulating materials and most fluids have low thermal
conductivities.
As an example, consider the flow of heat along a conductor whose cross-
section area is A = 0.1 in 2 and length 2 in, if the RMS current-density is
4,500 A /in2. The electrical resistivity of copper is 0.678 x 10 6 Ohm-in,
so heat is produced at the rate o f / 2p = 45002 x 0.678 x 10' 6 = 13.7
W /in3. In one conductor the / 2J?loss is therefore 13.7 x 0.1 x 2.0 = 2.75
W. To take the most pessimistic estimate, assume that all of this heat is
generated at the mid-point of the coil-side, half-way along the stack. The
thermal conductivity of copper is 9.8 (W /in2) per (C/in). So the
temperature gradient along the coil-side is given by equation (15.1) as
dT = _Q_ 2.75 2.8 C/in. (15.2)
dx kA 9.8 x 0.1
Since the heat can flow in both directions, the temperature-gradient is
only half this value, and the temperature rise between the ends of the
stack and the centre is therefore 2.8/2 x 2.0/2 = 1.4 C, which is
negligible. A more thorough analysis would have to consider the full
diffusion equation along the length of the coil-side, but this quick
calculation reveals that such sophistication is not needed in the example
considered.
15.2.2 Contact resistance
Equation (15.1) can be used to define thermal resistance as the ratio of
temperature difference A T to heat flow rate Q : the symbol used for
thermal resistance is R, with units C/W. Thus
(15.3)
The thermal resistance is a "lumped parameter" that can be used to
model the conduction through a region or interface where the individual
values of k, A, and t may be difficult to determine. The contact resistance
between two surfaces is usually treated in this way, as, for example,
between the frame and the stator core. The temperature drop across a
thermal resistance is given by equation (15.3) as A T~ QR For example,
if the contact resistance between the motor flange and the mounting

15-4
15. C o o l in g
plate is 1C/W, then with 40W flowing though it the temperature
difference across the interface would be 40 C.
The contact resistance between metallic surfaces held tighdy together
depends on the surface finish. For a 120pin milled finish the heat
transfer coefficient can be estimated as 0.7W/in 2/C , corresponding to
a thermal resistance of about 1.4C/W for an area of lin 2. A lapped
surface (5pin) might have a value of half this. Thermal grease can be
used to improve the heat transfer and lower the contact resistance by a
factor of approximately 2 by replacing the voids which otherwise would
be filled with air. The problem with thermal grease is that it tends to
migrate. An alternative is to use a gasket which may be made of
aluminium or copper foil, or special matrix materials impregnated with
graphite or silicone.

In electric machines, contact resistances are a major source of


uncertainty in thermal calculations, not only where the motor is
externally bolted to something else, but also internally, wherever heat is
required to flow across an interface. The most important interfaces are
those between the coil conductors and the slot insulation, between the
insulation and the lamination stack, and between the lamination stack
and the frame. It may be impossible to calculate them, in which case test
data is essential. The thermal contact resistances are usually the main
impedances to the flow of heat from the interior of the motor to the
outside, and consequently they make the largest contribution to the
temperature rise of the hottest internal parts.
For this reason it is common practice in servo motors to encapsulate the
windings in a resin of high thermal conductivity. Usually the thermal
conductivity of the encapsulation is much less than that of a metal, but
vastly better than that of air. In the most highly rated machines
(aerospace machines), direct cooling of the conductors is often necessary
because the overall thermal resistance between the windings and the
outside world is simply too high to permit an acceptable power density.

When conduction through the stator stack is the main path for heat
removal from the windings, it is advantageous to have a large number of
small slots in order to maximize the contact area between the core steel
and the windings and minimize the diffusion path length through the
slots for heat generated in the copper.

15-5
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

15.2.3 Radiation
Radiation is described by the Stefan-Boltzmann equation
S = e o i T f - T * ) W/in 2 (15.4)
A
where a is the Stefan-Boltzmann constant, 5.67 x 10 ~8 W /m^/K 4 or 3.66
x 10"11 W /in2/K for a black body. A black, body is a perfect radiator (no
reflection). Real surfaces are imperfect radiators and their effectiveness
relative to that of a black body is called the emissivity e. A black lacquered
surface can achieve an emissivity as high as 0.98, but a more practical
rule of thumb is to take 0.9 for black-painted or lacquered surfaces.
The temperature Tx is the absolute temperature of the radiating surface,
and T2 is the absolute temperature of the surroundings radiating back
to the motor .2 For example, a surface with an emissivity of 0.9 that is
50 C above the surroundings at 50 G, has a net heat transfer rate of
0.9 * 3.66x 10' 11 * ((50+50+273) - (50+273)4) (15-5)
which is 0.28 W /in2. A surface 30 C above the surroundings at 20G has
a rate of 0.12W/in 2 - quite a useful component of the heat-removal
capability of the frame.
15.2.4 Convection
Heat removal by convection is governed by Newtons Law.
-Q = Ji A T W/in2 (15.6)
A
where AT is the temperature difference between the cooling medium
and the surface being cooled, and h is the heat-transfer coefficient. The
units of h are W /in2/C [or W /m 2/C ]. The value of h depends on the
viscosity, thermal conductivity, specific heat, and other properties of the
coolant, and also on its velocity. In natural convection the flow of coolant
is not assisted by fans, blowers, pumps etc. In forced convection the flow is
assisted by one of these external means.

2 The absolute temperature in degrees Kelvin (K) is the temperature in C plus 273.
15-6
15. COOUNG
15-2.5 Natural convection
The heat transfer coefficient for natural convection around a
horizontally-mounted unfinned cylindrical motor can be roughly
estimated as
h (15.7)
For example, for an unfinned cylinder of diameter 4 in and a tempera
ture rise of 50 C, the natural-convection heat-transfer coefficient is
calculated as 0.0040 W /in 2/ 0C. For a A T of 30 C the heat transfer rate
is then 0.12W/in2. As a first approximation this value can be applied to
the whole surface including the ends, but if the motor is flange-mounted
then only one end is available for convective cooling.
15.2.6 Forced convection
Forced convection, with "air-over" cooling from a shaft-mounted or
external fan, increases the heat-transfer coefficient by as much as 5-6
times, depending on the air velocity. The increase in heat-transfer
coefficient is approximately proportional to the square-root of the air
velocity. An approximate formula for the forced-convection heat-transfer
coefficient is
h - 11.2 * 10^ (15.8)
where V is the air velocity [ft/min] and L is the frame length [in]
(assumed parallel to the direction of airflow). For a motor of length 3.7
in, if the air velocity is 800 ft/min, this formula predicts h = 0.0165
W/in 2/C . This is 4 times higher than for natural convection.
The air velocity V is the actual air velocity, not the so-called "no-load"
value.The no-load flow through a fan is usually specified in cubic
feet/min (CFM), and the no-load velocity is given by the no-load CFM
dividedby the fan inlet area (in ft2). The actual air velocity is determined
from the intersection of the curve of static pressure vs. flow rate for the
fan,and the pressure/flow curve for the air path over the motor. This
calculation requires the use of fluid-dynamics, but a rough guide is to
takeV as one-half the no-load value:

15-7
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

y _ No-load CFM ft/mjn (15 9)


2 * Fan inlet area
The size of fan required can be roughly estimated from the formula
CFM - (15.10)
^ 'air

where Qis the total rate of heat removal and A is the temperature-
rise of the air passing over the motor. Normally A T^t should be limited
to about 15C for an ambient temperature of 50 C. For example, if 100
W is to be removed with a A Tair of 15C, the actual CFM of the fan must
be 11.7 CFM. To allow for static pressure drop, a no-load fan of 25-30
CFM should be considered.
15.2.7 Some rules of thumb for "calibration"
Holman [1] gives an interesting example of a water-immersed wire 1 m
long, 1 mm diameter, in which a power loss of 22 W (0.56 W per inch
length) is sufFicient to boil the water at the wire surface. The wire surface
temperature is 114C and the heat transfer coefficient (see below) is
5000 W /m 2/C or 3.23 W /in 2/C . The heat flow at the wire surface is
45 W /in 2 (0.07 W /mm2) and the current-density in the wire is
approximately 35 A/mm2.
In normal motors, the rate at which heat can be abstracted is nothing like
as high as this. Correspondingly, current-densities as high as 35 A/mm 2
are achievable only for very short bursts. This current density is sufficient
to fuse a copper wire in free air.
The maximum rate at which heat can be removed from a surface by
natural convection and radiation (with 40 C rise) is only about 0.5
W /in2. With forced air convection the rate increases to about 2 W /in2j
and with direct liquid cooling about 4 W /in2. Motors that generate more
heat than can be removed at these rates have to absorb the heat
internally in their thermal mass, which is an acceptable way of increasing
the output power for a short time.
These rates limit the heat generated per unit volume to about 0.2 W/in
for natural convection, 5 W /in 3 for4 metallic conduction, 7 W /in 3 for
forced-air convection, and 10 W /in for direct liquid cooling.

15-8
15. C o o l in g
If rated torque is required at very low speed, a shaft-mounted fan may not
provide enough coolant flow to keep the motor cool. DC motors often
have separate AC-driven fans, because they have to work for prolonged
periods at low speed with high torque. Since most of the heat in a DC
motor is generated on the rotor (in the armature windings and the
commutator), a good internal airflow is essential for cooling. In DC
motors the external fan is usually mounted to one side of the motor,
where it is easily accessible, and does not increase the overall length. A
similar problem arises with AC induction motors, especially vector-
controlled motors. A common practice is to mount the fan in line with
die motor at the non-drive end, and arrange it to blow air over the
outside of the finned frame. The fan may increase the overall length by
as much as 60%. Brushless motors do not have this problem to the same
degree, because most of the heat at low speed is generated in the stator
windings, where all three forms of cooling (conduction, radiation and
convection) are more effective.

15.3 Internal temperature distribution


The heat-removal calculation determines whether a steady-state case
temperature can be achieved for a given value of total losses. This is
important in choosing the method of cooling, bat it tells nothing about
die internal temperature distribution, and gives no guidance on the
values of current-density, flux-density, and frequency that can be used.
15.3.1 The diffusion equation
The internal temperature distribution is the result of heat-flow processes
within the motor, including conduction, radiation, and convection. For
conduction alone, the partial differential equation describing the three-
dimensional flow of heat is the so-called diffusion equation:
V*:T+ - 1 i? (15.11)
k dt a dt
where
v l r , ? T t ? T t #T (] 5 i 2)
d* 2 t y 1 dz 1
and

15-9
D e s ig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

= J L m 2/s (15.13)
pc
n
is the dijfustvity in SI units. (English units are in /s). In SI units, k is the
thermal conductivity in W/mC; cis the specific heat in kJ/kgC, and p
is the density in kg/m 3. In a structure as complex as an electric motor
the heat conduction equation is a complex boundary-value problem that
is best solved by computer-based numerical methods such as the finite-
element method.
In electric motors the internal convection and radiation processes may
be as important as the conduction process, and when the differential
equation is extended to include them, matters become very complicated.
15.3.2 Thermal equivalent circuit
For most practical purposes it is sufficient to use a thermal equivalent circuit
of the interior of the motor, Fig. 15.1. The thermal equivalent circuit is
an analogy of an electric circuit, in which heat is generated by "current
sources" and temperature is analogous to voltage. The rate of generation
of heat in a source is measured in Watts. The heat flow rate, which is also
measured in Watts, is analogous to current. Resistance is measured in
C/W. The copper losses, core losses, and windage 8c friction losses are
represented by individual current sources, and the thermal resistances of
the laminations, insulation, frame, etc. are represented as resistances. In
the simplest possible model, all the losses are represented together as
one total source, i.e. the individual sources are taken as being in parallel.
Heat source W inding H ot Spot
(losses) T S

Am bient tem perature


Fig. 15.1 Thermal equivalent circuit

15-10
15. C o o l in g
The thermal equivalent circuit is really a lumped-parameter model of all
the heat-flow processes within the motor as well as the heat removal
processes discussed in section 15.2.

A simple thermal equivalent circuit can be constructed relatively easily,


ft should ideally take into account the anisotropy effects: for example,
the effective thermal conductivity through a lamination stack is lower in
the axial than the radial direction.

Pd
0

Fig. 15.2 Temperature variation caused by intermittent operation

The thermal equivalent circuit shown in Fig. 15.4 includes provision for
direct cooling of the winding conductors, ("cad" = eonductors-to-
a m b ie n t/ direct); and for direct cooling of the rotor shaft ("had" =
staft-to- ambient direct). It also includes the thermal mass or capacity of
the winding , and the thermal capacities of the rotor and stator
laminations, and The other internal thermal resistances are
essentially self-explanatory.
The heat removal routes by conduction, radiation, and convection are
also represented by thermal resistances. For convection the appropriate
resistance is given by
15-11
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

^
= J - -C/W (15.14)
hA
where A is the appropriate surface area for convective heat-transfer and
the subscript "v" stands for convection. If A is a function of the
temperature-difference, the equivalent circuit becomes non-linear and
requires an iterative solution. For radiation the equivalent thermal
resistance is the ratio of the temperature difference 1\ - T2 to the
radiation heat exchange rate Q in equation (15.4). Clearly this is non
linear. However, the non-linearity is often neglected and a fixed value of
Rf is calculated assuming that the final temperature of the case is known.
15.3.3 Current Density
The current-density cannot be directly related to the temperature rise of
the winding by a simple general equation, because the heat transfer rate
depends on the shape of the conductors. For example, 1 in3 of copper
can be made into a stubby cylinder of 1 in diameter and 1.27 in length,
or a long wire of 0.5 mm diameter and 83 m length. If only the
cylindrical surface area is available for cooling, the short cylinder has a
surface area of 4 in2 while the lone wire has a surface area of .2039in2.
The loss density in W /in o in copper conductor is J n p where yis in A/in
and p is in p-in. At 20 C the resistivity of copper is approximately 0.68
pQ-in, but at 100 C it is about 1.16 p2-in. If 1 W can be dissipated from
every in2 of surface at 100C, this suggests that in the short stubby
cylinder the permissible current density is 2,115 A /in2 (3.3 A/mm 2) and
in the long wire, 15,076 A /in2 (23.4 A/mm 2).
With this reservation, it is still possible to quote typical values of current
densities used in motors cooled by different methods:

Cooling method A/in2 A/mm2


TENV 3000 - 3500 4.7-5.4
Air over; fan-cooled 5000 - 7000 7.8 - 10.9
External blower; through-cooled 9000- 10000 14.0 - 15.5
Liquid-cooled 15000- 20000 23.3 - 31.0
T able 15.1 T ypical values o f current -density

15-12
15. C o o l in g

Motor type Class B Class F Class H


1.15 Service Factor 90 115 140
1.00 Service Factor 85 110 135
TEFC 80 105 125
TENV 85 110 135
(NEMA Standard MG-1), "C. Assumes 40 C ambient temperature.
T a b le 15.2 T em pera tu re r ise by r e sist a n c e a n d in s u l a t io n

Material p <20C) k Sp. Heat Density


Q-m x 10-8 (W/m K) kJA g/'C kg/m
Copper 1.72 360 0.38 8950
Aluminium 2.8 220 0.90 2700
0.1% Carbon steel 14 52 0.45 7850
Silicon steel 30-50 20-30 0.49 7700
Cast iron 66 45 0.5 7900
Cobalt-iron 40 30 0.42 8000
Ceramic magnet 104 4.5 0.8 4900
Re-Co magnet 50 10 0.37 8300
NdFeB magnet 160 9 0.42 7400
Kapton$ 303 V/pm* 0.12 1.1 1420
Teflon 260V/pm* 0.20 1.2 2150
Pressboard/Nomex 10kV/0.22mm* 0.13 1000
Epoxy resin SOkV/mm* 0.5 1.7 1400
Water (20C) 0.0153 4.18 997.4 I
Freon 0.0019 0.966 1330 1
Ethylene Glycol 0.0063 2.38 n 17 |j
Engine oil 0.0037 1.88 888 |

T able 15.3 S elected material properties *Dieiectric strength


15-13
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

Material Emissivity
Polished aluminium 0.04
Polished copper 0.025
Mild steel 0.2-0.3
Grey iron 0.3
Stainless steel 0.5-0.6
Black lacquer 0.9-0.95
Aluminium paint 0.5
T a b le 15.4 Selected e m issiv it ies

15.4 Intermittent operation3


Intermittent operation is normal for brushless PM motors, because most
of the applications that use them are motion-control applications with
programmed moves, acceleradons, deceleradons, stops, starts, and so on.
Consequendy the temperatures of the windings and magnets are
constandy varying. A simple example is shown in Fig. 15.2, where the
motor executes a simple on-off sequence: on for <qN and off for t^ ^ ,
after which the on/off cycle repeats indefinitely. The cycle time (cy is
'e y = 'on + W

15.4.1 Duty-cycle
The duty-cycle d is defined as
d = . (15.16)
'e y ' o n + 'o f f

The most efficient use of the thermal capability of the motor will be
9
This section is an expanded analysis based on ideas originally published in the paper
"Dynamic thermal model for a three-phase sinusoidal Ag brushless servomotor" by Richard
Welch and George Kaufman of Reliance Motion Control.
15-14
15. C o o l in g
made if the maximum winding temperature T just reaches the rated
value Tt at the end of each on-time. Because the power dissipation is
interrupted with cool-down intervals (OFF, the power that can be
dissipated during the on-times may exceed the steady-state continuous
dissipation rating of the motor PT, and therefore the motor may be
permitted to exceed its steady-state output power rating during the on-
times. The simplified thermal equivalent circuit model in Fig. 15.1 makes
it possible to calculate the permissible overload factor as a function of
the on-time <ON and duty-cycle d for a given motor.
The thermal equivalent circuit is a parallel combination of thermal
resistance R and thermal capacitance C. R represents the steady-state
thermal resistance between the winding and the surroundings in C/W.
C represents the thermal capacity of the entire motor in J/C . The
thermal time-constant t is given by
r = RC (15-17)
in [s].
The analysis is based on equating the temperature rise during the on-
time with the temperature fall during the off-time. To do this we need
the equations for the temperature rise and the temperature fall.
15.4.2 Temperature rise during ON-time
During the on-time (qN> the power dissipation in the motor is Pd and the
temperature rises according to the equation
T - T0 = RPd( 1 - e-fr) + (Tc - T0)e -tlT. (15.18)

The temperature rise is expressed relative to the ambienttemperature


T0. The second term in equation (15.18) is due to the initial condition
in whichthe temperature rise is (Tc - T0) at t = 0. At t = tQN>

- To = " e ' WT) + ( r c - TJa'***.( 15

15-15
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

Overload fa c to r k

Relationship between duty-cycle, on-time thermal time-constant, and


overload factor for intermittent operation in which the maximum motor
temperature reaches 7J, momentarily in each cycle
15. C o o l in g
By definition, the steady-state rated temperature-rise (Tj. - T0) is given by
Tt - T0 = RPt , (15.20)
where Pr is the rated steady-state power dissipadon in the motor, i.e., the
continuous power dissipadon that produces rated temperature rise. We
can use this to "calibrate" PA in equation (15.18), by defining the
dissipation overload factor k2, where
it2 = (15.21)
The reason for using A2 instead of k is th a t in most types of brushless
servomotor the losses are dominated by p R losses while the load torque
is proportional to the current I. If the load is increased by a factor k, it
means that the current and torque are increased by the factor h while the
losses increase by t?. Thus k is the overload factor for torque and
current.
Substituting equations (15.20) and (15.21) in equation (15.19) and
rearranging, and assuming that
Trcax = Tt , (15-22)
we obtain the following equation relating the temperature rise to the
overload factor k and the on-time *qN:

(Tt - 7J)[1 - P (1 - e 'W r)] = (Tc - 7J)<fW r (15.23)

15.4.3 Temperature fall during OFF-time


When the motor is switched off, the power dissipation falls to zero and
the winding temperature falls according to the equation
T - T0 = ( Tt - T0) e 't>T (15.24)
where t is measured from the end of the on-time, i.e.the beginning o
the off-time. At
T c - T0 = (7 ; - 7 J ) e 'W r . (15.25)

15-17
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

15.4.4 Steady-state : Equating the temperature rise and the temperature fall
First, multiply equation (11) by e rN/r:
( Tc - T J e '0H'T = ( Tr - T Je<to* * W)/r. (15.26)
The left-hand side of equation (15.26) is identical to the right-hand side
of equation (15.23), so the right-hand side of equation (15.26) can be
equated to the left-hand side of equation (15.23). With suitable
rearrangement, the result can be expressed in different ways, all of which
are useful for different purposes.
15.4.5 Maximum overbad factor
First, we get a solution for the dissipation overload factor A2 in terms of
the on-time and the duty-cycle: writing <q N/<2 instead of <q N + i.e.,
instead of t^, the expression is
P = 1" g - (15.27)
1 -

For example, if the duty-cycle is 25% (d = 0.25) and <ON = 0.2 x t, the
dissipation overload factor is
k2 = L l = 3.04, (15.28)
1-
which means that the dissipation can be increased to 304% of its rated
steady-state value for a period of /qN = 0.2t in every cycle of length t^ =
<QN/d = (0.2/0.25)t = 0.8t. A s a concrete example, if t = 40 min, then
the dissipation can be raised to 304% for 8 minutes followed by a cool
down period of 24 minutes.
In the above example, increasing the dissipation to 304% of the rated
value corresponds to an increase in current and torque to ^3.04 =
1.74 times iheir rated values, or 174%.
If t t, then equation (15.27) simplifies so that
k 2 = -i. (15.29)
d
15-18
15. C o o l in g
This means that when the on/off cycles are very short compared with the
thermal time-constant of the motor, the mean dissipation will be equal
to Pr when the peak dissipation i j = f?Pr is equal to Pr/d . This simple
result, is intuitive.
15.4.6 Maximum overload for a single pulse
Equation (15.27) can also be used to calculate the maximum dissipation
overload factor fora single pulse, for which d = 0. In this case
Jc2 = -1 ~. (15.30)
1 - e 'W

For example, if CO N = 8 min and t = 40 min, then the maximum


dissipation overload factor A2 is 5.5 or 550%, and the maximum overload
factor k is 2.35 or 235%.
15.4.7 Required cool-down period for a given overload factor and on-time.
The second result that arises from equating the temperature rise and
temperature fall is an expression for the necessary cool-down time iOFF
as a function of the dissipation factor &2 and the on-time /q N. The
expression is
' o f f = " T 111 t* 2 ~ (* 2 - 1 ) 0 ^ ] . (15.31)
Together with equation (15.16), this can be used to determine the
maximum duty-cycle d that can be used with a given dissipation overload
factor I? and a given on-time (qN, for a motor of thermal time-constant
T. For example, if the dissipation is 200% of rated, and if = 8 min,
t = 40 min,
' off = ' 40 * lnt2 - (2 1) * e8/4l = 100 min. (15.32)
The minimum cycle time is therefore 18 min and the maximum duty-
cycle (with 8 minutes on-time) is 8/18 = 0.44 or 44%.
15.4.8 Maximum on-time for a given overload factor and cool-down time
A third result arising from equating the temperature rise and fall is an
expression for the maximum on-time /qN as a function of the dissipation
overload factor k2 and the off-time *OFF. The expression is
15-19
D esig n o f b r u sh l e s s pe r m a n e n t -m a g n e t m o t o r s

toti = r In (15.33)

15.4.9 Maximum duration of single pulse


This expression can be used to calculate the maximum duration of a
single pulse having a given dissipation overload factor k2. For a single
pulse, jQpy is infinite and
(15.34)

For example, if k2 - 5.5 and t = 40 min, then /QN = 8 min.


15.4.10 Graphical transient heating curves
Fig. 15.3 shows the relationship expressed by equation (15.33) graphically
in terms of the duty-cyde d, the on-time (qN as a fraction of the time-
constant t , and the overload factor k.
This graph can be used in a number of ways. For example, to determine
the maximum permissible duration of a single pulse with a given
overload factor k, the duty-cycle d should be set to zero. Thus with k= 1.5
the maximum pulse duration is 0.58t. With a time-constant of 40 min
this is 23.2 min.
The graph shows the maximum duty-cycle that can be used with a given
overload factor. For example, at 200% load the maximum duty-cycle is
0.25 or 25%, but in this limiting case the on-time must be vanishingly
small. With an on-time of 0.1 T at 200% load, the maximum duty-cycle is
approximately 0.2, which means that the cool-down period in each cycle
must be (1-<2)t = 0.9t. If x is 40 min, this means a maximum operating
time at 200% load of 4 min, followed by a cool-down period of 36 min
before the cycle can be repeated. Operations that require a short on-time
with a high duty-cycle must use a lower overload factor.

15-20
15. COOLING

15-21
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

Thermal Units
capacitance kJ/C
Thermal Units
resistance C/W

Heat Units
source W

RcsisUnee: from to by
Frame Ambient conVeclion
Conductor! conduction
Rotor conVection
Si at or Radiation
lHafl Direct
Ambient Total
Fig. 15.5 Key for Fig. 15.4

15.5 Thermal modelling by computer


15.5.1 Computer model of thermal equivalent circuit
Fig. 15.3 is useful as a guide to the relationships between on-time, duty-
cycle, thermal time constant, and temperature rise, but it is based on the
assumption of regular variations. In many applications the variations are
quite irregular and computer simulation is then necessary to calculate the
thermal response in any detail. Fig. 15.4 shows an example of a detailed
thermal equivalent circuit model used for this purpose.
The computer program associated with this model integrates the
differential "circuit" equations so that transient temperature-vs-time
curves can be plotted, showing the initial rate of rise and the final steady
temperature at several points in the motor cross-section. Such a
computer program is invaluable for identifying where the bottlenecks are
in heat transfer, and for determining the permissible short-time overload
with due regard for the distribution of temperature inside the motor.
15-22
15. C o o l in g
Because the thermal equivalent circuit model in Fig. 15.4 is generic, the
computer code can be easily incorporated in other design programs such
as PC-BDC (see Chapter 12) and its stablemates.
15.5.2 Determination of thermal equtvalent-circuit parameters by test
Earlier sections in this chapter discussed the calculation of thermal
resistance from physical principles, but some of the interface resistances
and diffusion resistances are extremely difficult to calculate. Examples
are the thermal interface resistance between the coilside in a slot and the
surrounding steel, or the diffusion resistance through the coilside itself,
which is often a compacted jumble of insulated copper wires with no
particularly regular shape. It is clearly impossible to calculate precisely
the thermal resistance between adjoining conductors.

Eg. 15.6 Simple apparatus Tor measuring heat transfer of conductors in slots
In these cases experimental data is essential, and Fig. 15.6 shows an
example of a special test fixture for measuring the heat transfer between
a conductor in a slot, and the surrounding stator. The block is made of
Aluminium, which has a high thermal conductivity so that it can be
assumed to be at a uniform temperature throughout its cross-section.
(This is verified by test). The block used in the SPEED Laboratory is a
15-23
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a c n e t m o t o r s

3-inch cube with a slot width of 10mm and a slot depth of 30mm.
Thermocouples are let into the block and into the compression yoke at
the locations shown in Fig. 15.6 by small circles.
The slot can be lined with various slot-liner materials, and then a copper
coil is laid in the slot. The coil is wound on a former sized such that its
length overhangs the ends of the slot by the minimum manageable
am ount Typically 75% of the copper lies within the slot and 25%
outside. The coil may be varnished or potted, and it can be compressed
to a predetermined and repeatable pressure by a weight on top of the
compression yoke. The yoke is insulated from the coil by an insulating
shoe. Coils of varying depth/width ratio can be laid in the slot, and a
thermocouple is laid in the middle of the coil to meaure the hot-spot
temperature.
The coil is excited by DC to raise its temperature, and transient
temperature curves are plotted for all the thermocouples. By varying the
shape and size of the coil, the slot liner material and thickness, and the
degree of compression, it is possible to build up an extensive database of
heat-transfer data which can be used in the thermal model of Fig. 15.4.
It is possible to study the diffusion of heat through the coil by laying two
or more coils on top of each other, and exciting only one of them.
Thermocouples are provided in the middle of each coil.
The heat source in Watts is easily calculated by measuring the DC voltage
and current, and then multiplying the power by the fraction of copper
(by weight) that is in the slot- This fraction is accurately determined by
cutting off the end windings at the end of the experiment, and weighing
the bits.
The heat transfer block can be used to measure the thermal conductivity
and specific heat of insulating materials. In this case a suitable thickness
(e.g. multiple layers) of an insulating material is laid at the bottom of the
slot, and an Aluminium shoe with an embedded thermocouple is laid on
top. The heating coil is laid on top of the shoe. This provides a regular
rectangular geometry from which the thermal conductivity and specific
heat can be extracted following the thermal experiments.
Reference
1. Holman JP [1989] Heat Transfer, McGraw-Hill ISBN 0-07-100487-4

15-24
16. MAGNETIC MATERIALS
16.1 Introduction
This chapter reviews the material properties of permanent magnets,
electrical steel and insulated wire.
There is still widespread use of c.g.s. units in the magnet industry,
whereas motors are usually designed in metric (SI) units in Europe and
Japan, and in metric or mixed units in the USA. The most important
magnetic conversion factors are given in Chapter 4.

16.2 Permanent magnets


A brief description of the characteristics, parameters and equations
associated with permanent magnets is given here, i.e. the normal and
intrinsic hysteresis curves, major and minor loops, temperature effects,
energy product, magnetization, etc. More detailed descriptions can be
found in specialist works on permanent magnets, e.g. [1].
16.2.1 The hysteresis loop and demagnetization characteristic
Fig. 16.1 shows a typical hysteresis loop in both normal and intrinsic
forms. The normal curve shows the total flux density BMas a function of
the externally applied magnetic field strength HM and the intrinsic
magnetization M of the magnet material. The intrinsic curve shows the
intrinsic polarization, J . These parameters are related as follows:
~ 4J (16.1)
J = \Lq M (16.2)
~ Mo + (16.3)
The important points on the characteristic are:
1. Remanence B the value of flux density corresponding to zero
externa] applied field, HM = 0. It corresponds to a "magnetic
short circuit", which would be obtained if the magnet was
surrounded by an infinitely permeable "keeper".
16-1
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

Fig. 16.1 Typical hysteresis loop of 'hard" permanent magnet material.

2. Coercivity H the value of magnetizing force that must be applied


to reduce the flux density to zero: i.e., the value of when B^
= 0.

3. Intrinsic coercivity Hc the value of magnetizing force that must be


applied to reduce the intrinsic polarization to zero: i.e., the value
of H m when / = 0.
4. Relative recoilpermeability Hrecthe gradient of the B /H curve at the
remanence point, relative to fig. The relative recoil permeability
of "hard" magnets is in the range 1.0-1.1, i.e. close to that of air.
It may be surprising that magnets have low permeability, but this
is an important property because it helps to limit the
demagnetizing effect of armature reaction (see Chapter 12).

16-2
16. M a g n e t ic m a teria ls

5. Limiting or knee-potrU magnetizing force H^the value of //M in the


second quadrant at which the //WM curve starts to become non
linear. If the magnet working point is operated to the left of
there will be a degree of irreversible demagnetization. This can
only be recovered by re-magnetization.
Magnet materials come in both isotropic and anisotropic forms. An
isotropic material has the same properties in all directions. Anisotropic
materials have a preferred direction of magnetization with increased
remanence and coercivity.
Fig. 16.2 shows the 1st and 2nd quadrant curves of a typical B/H
hysteresis loop. The material is initially unmagnetized at point A. First,
an external magnetizing force (//M > 0) is used to drive the working
point along the "initial magnetization curve" to point B. If the externally
applied magnetizing force is removed, the working point will move to
point C on the major curve in the 2nd quadrant. The position of point

16-3
D e s ig n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

C depends upon the shape of the magnet and the working permeance
of the circuit it is supplying and can be calculated using a load line*
technique (Chapter 4). The 2nd quadrant of the major B-H loop is
referred to as the "demagnetization curve". If the applied magnetizing
force is insufficient to magnetize the magnet fully, it will work on a
minor loop inside the major loop.
The working point may be driven further down the B-H curve by
applying an external demagnetizing field (//M < 0). As long as the
working point is not driven below the knee D, the reduction in B ^ is
reversible: i.e., when the external field is removed, the working point will
retrace the demagnetization characteristic back to point C. If, on the
other hand, the working point is driven below D, say to point E, then
when the external field is removed the working point does not retrace
the demagnetization curve, but follows a recoil line at a lower level of
flux-density. The recoil line may be considered a straight line (E-F), of
constant slope equal to the recoil permeability (-Lo*J-rcc
If the working point is driven below the knee D, the magnet will be
partially irreversibly demagnetized. Having been exposed to this sequence
of operations, the magnet is stabilised against further irreversible
demagnetization by external fields smaller than the maximum negative
value of Hm.
16.2.2 Permanent magnet materials
Much of the recent progress in the development of permanent magnet
brushless motors can be attributed to remarkable improvements in the
properties of magnet materials. Fig. 16.3 shows that since 1900 the
historical development in maximum energy product of commercial
permanent magnets has been nearly exponential. Fig. 16.4 compares
typical B /H characteristics of the major magnet material types. A brief
summary of magnet properties is also given in Table 16.1. More detail
can be obtained from suppliers data sheets.
The metal alloy materials are seldom used in motors where the magnet
is placed adjacent to the airgap, since their highly non-linear
demagnetization characteristics and low coercivities give them a low
resistance to demagnetization. However, the essentially linear
demagnetization characteristics and relatively high coercivities of ferrite
and rare-earth magnets make them ideally suited to motor applications.
16-4
16. M a g n e t ic m a teria ls

YEAR
Fig, 16.3 Improvement! in (BH)mll since 1900

There has been continuous development of ferrite materials since their


introduction in 1953. They are nearing their theoretical maximum
properties and are well established as the most cost-effective material for
many types of motor. Rare-earth Samarium-Cobalt (Sm-Co) was
introduced in the mid 1970s. It has a much larger energy product than
ferrite and is also thermally very stable. It is however very expensive and
thus will continue to find use only in the most technically demanding
applications requiring precise performance over extended temperature
ranges.

16-5
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

Properly Units Alnico Anisotropic Sintered Sintered


Ferrite Sm-Co Nd-Fe-B
Remanence BT T 0.6 to 1.35 0.35 to 0.43 0.7 to 1.05 1.0 to 1.3
Intrinsic kA/m 40 to 130 180 to 400 800 to 1500 800 to
Coercivity 1900
Hd
Recoil 1.9 to 7 1.05 to 1.15 1.02 to 1.07 1.04 to 1.1
Permeability
Mr
kj/m5 20 to 100 24 to 36 140 to 220 180 to 320
1.
j

Magnetizing kA/m 200 to 600 600 to 1700 1600 to 2000 to


force 4000 3000
Resistivity pQcm 47 >104 86 150
Thermal lOV C 11.3 13 9 3.4
expansion
%/C -0.01 to -0.2 -0.045 to -0,08 to
tem perature -0.02 -0.05 -0,15
coefficient
Hci % rc -0.02 0.2 to 0.4 -0.2 to -0.5 to
tem perature -0.25 -0.9
coefficient
Max. working c 500 to 550 250 250 to 350 80 to 200
temperature
Curie *c 850 450 700 to 800 310 to 350
temperature
Density kg/ms 7300 4900 8200 7400

T able 16.1 Magnet properties (20G)

16-6
16. M a g n e t ic m a t er ia l s
1.40

(a) ANI SOTROPIC ALNICO <aV/


b) ISOTROPIC ALNICO 1.20
c) SINTERED ANISOTROPIC FERRITE
d) POLYMER BONDED Sm-Co >
a) POLYMER BONDED Nd-Fe-B / <J
1.0 0 -J'
(f) SINTERED Sm-Co X W
(g) SINTERED Nd-Fe-B 0.80 pj
/
/ yT J / X 0.60 CO
<,oy y / yf /
/
0.40

y' y / (B) / \s// 0.20

-900
k i
-800 -700
--i
-600
,/T T i M
-300 -400 -300 -200 -100 0
0
H (kAAn)
Fig. lfi.4 Typical B/II curves
A new generation of rare-earth Neodymium-Iron-Boron (Nd-Fe-B)
materials emerged in the mid 1980s. Nd-Fe-B has a larger energy-density
than Sm-Co and is made from less expensive materials. However, Nd-Fe-B
is by no means as thermally stable as Sm-Co and suffers from corrosion
problems. Nd-Fe-B was pioneered by Sumitomo in Japan and General
Motors in the USA. Both use similar compositions, but use fundamentally
different processing techniques for production. The initial predictions
that Nd-Fc-B materials would displace Ferrite in many applications has
not yet materialised. This is mainly due to the fact that Nd-Fe-B is more
costly than ferrite and that current grades of Nd-Fe-B have a limited
operating temperature range and are relatively susceptible to corrosion.
These factors have probably helped ferrite to remain the magnet of
choice for most automotive applications, and rare-earth Sm-Co for most
aerospace applications and high-performance servomotors. However, Nd-
Fe-B is making good progress and is now used in many motor products.
The corrosion can be controlled by suitable coatings (or in some cases
by encapsulation or even the use of hermetic canisters as in the
celebrated GM cranking motor). Considerable research worldwide is
devoted to advancing the technical capabilities and application of Nd-Fe-
B materials.

16-7
D e s ic n o f b r u s h l e s s pe r m a n e n t -m a g n e t m o t o r s

16.2.3 Temperature effects


The characteristics of all permanent magnets vary with temperature to a
greater or lesser degree. The effect of temperature-induced changes is
represented by the temperature coefficients of remanence ( t Br) and
intrinsic coercivity ( T H c i ) . Typical values are given in Table 16.1. Ferrite
and Nd-Fe-B magnets are particularly susceptible to changes in
temperature. However, as the ferrite material has a positive value of
over most of the temperature range of interest in motors, and vice-versa
for Nd-Fe-B material, ferrite material has a tendency to undergo
irreversible demagnetization at its lowest temperature (Fig. 13.2), and Nd-
Fe-B at its highest temperature, (Fig. 13.5). For ferrite material, it is
possible to predict the J-H characteristic at any temperature from the
ambient J-H characteristic and the essentially constant values of Tgr and
THcj. However, this is not usually possible for Nd-Fe-B since THci is usually
a non-linear function of temperature.
Some of the highest-temperature applications are in aerospace, for
example PM generators on aircraft engines. "2-17" Sm-Co is now widely
used in high-temperature applications up to 250 C, but Alnico magnets
can be used at even higher temperatures. "2-17" is also widely used in
industrial servo motors because of the need for a high torque per unit
volume, high torque per ampere, and good thermal stability in a totally
enclosed machine where cooling is limited and high temperatures may
be encountered.
Exposure to high temperatures for long periods may produce
metallurgical changes which may impair the ability of the material to be
magnetized and may even render it nonmagnetic. There is also a
temperature, called the Curie temperature, at which all magnetization is
reduced to zero. After a magnet has been raised above the Curie
temperature it can be remagnetized to its prior condition provided that
no metallurgical changes have taken place.
The temperature at which significant metallurgical changes begin is lower
than the Curie temperature in the case of the Sm-Co, Nd-Fe-B, and
Alnico; but in ferrite magnets it is the other way round. Therefore ferrite
magnets can be safely demagnetized by heating them to a temperature
just above the Curie point for a short time. This is useful if it is required
to demagnetize them for handling or finishing purposes.

16-8
16. M a g n e t ic m a t er ia l s
Table 16.1 gives typical values of Curie temperature and maximum
recommended working temperature for sintered magnet materials. The
maximum recommended working temperature for all the materials is
sufficiently high for most practical motors. However, if the magnets are
in their bonded forms, there is an upper temperature limit due to the
properties of the polymer binder material. This is an area in which there
has been much recent research and the latest bonded materials have a
temperature limit of around 180 C.
16.2.4 Magnet energy product
It is shown in Chapter 4 that for a given airgap volume magnetized to a
certain flux-density, the required magnet volume is inversely proportional
to the energy product, i.e., the product B ^H ^. Contours of constant energy
product are rectangular hyperbolas, frequently drawn on property data
sheets provided by magnet suppliers. The maximum energy product or
(BH)m3X of a given magnet occurs where the demagnetization
characteristic is tangent to the BH hyperbola. If the recoil permeability
is unity, this occurs for a permeance coefficient of unity, provided that
there is no externally applied field from windings or other magnets.
In static magnetic circuit designs where there is no demagnetizing MMF
from electric currents, the magnet length and pole area can be
proportioned relative to the length and area of the airspace, so as to
cause the magnet to work at (BH)m3X. In motors this principle must not
be followed because the armature current produces demagnetizing
ampere-turns that may be very great under fault conditions (Chapter 12).
To reduce the risk of demagnetization, motors are designed so that on
open-circuit or no-load, the magnet operates at a high permeance
coefficient (corresponding to a small airgap length) with adequate
margin of coercive force to resist the maximum demagnetizing
ampere-turns expected under load or fault conditions. Some designs rely
on the current-limit in the controller to guard agains demagnetization.
The B fly product is sometimes preferred over (BH)m as a figure of
merit of magnet materials for motors, especially when the magnets are
exposed to potentially large demagnetizing fields, as in surface PM
motors. The /?r//k product is simply the value of remanence (Br)
multiplied by the knee value of magnetizing force (i/k). This product is
a combined measure of the flux-producing capacity with the resistance
to demagnetization under the influence of external MMF.
16-9
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

16.2.5 Magnetization
The value of magnetizing force H required to "saturate" a magnet varies
with the type of material. The saturation magnetizing force can be taken
approximately as a multiple k of the intrinsic coercivity, where k is about
3 for anisotropic and 5 for isotropic materials. If a lower value of A is
used, the magnetization will be less than the saturation magnetization,
which is the maximum magnetization achievable that produces the
maximum remanence.
Table 16.1 lists typical values of magnetizing force requirement for some
common magnets. High-energy magnets require such a high magnetizing
force as to need special fixtures and power supplies, and this is one
reason why high-energy magnets are usually magnetized before shipping.
Even then, the ampere-tums requirement is usually beyond the
steady-state thermal capability of copper coils. Therefore, pulse
techniques are used, or in some cases superconducting coils.
Ceramic (ferrite) and Alnico magnets can sometimes be magnetized in
situ in the final assembly, but this is seldom possible with high-energy
magnets, particularly in motors because of the airgap, the slotting, the
circular geometry (which increases "leakage" flux during magnetization),
the limited conductor cross-section, and the generally inappropriate
disposition of conducting coils. Finite-element analysis is often useful for
designing the magnetizing fixture and to determine the magnitude of the
applied field required to give complete magnetization. In some cases is
it impossible to achieve complete magnetization: the demagnetization
characteristic can then be inferior to the major characteristic [8,10] and
this must be taken into account in the motor design.
In some applications bonded magnets are preferred over sintered
magnets because it is relatively easy to form the magnets into complex
shapes which can then be magnetized with any pole configuration.
16.2.6 Mechanical properties and handling
Magnets are often brittle and prone to chipping, but proper handling
procedures are straightforward enough as long as the rules are followed.
Modern high-energy magnets are usually shipped in the magnetized
condition, and care must be taken in handling to avoid injuryfor

16-10
16. M a g n e t ic m a teria ls

example, trapped fingers or particles getting in the eyes. A further


hazard is that when two or more magnets are brought close together they
may flip and jump, with consequent risk to eyes. Table 16.5 summarizes
some of the important safety precautions.
The best way to "tame" magnetized magnets is to keeper them. Fixtures
for inserting magnets can be designed so that the magnets slide along
between steel guides which are magnetically short-circuited together.
There sdll remains the problem of entering the magnets between the
guides, but usually there is enough space to provide for this to be done
gently. Fig. 16.5 shows this technique used for inserting magnets in an
interior-magnet motor rotor.1
Obviously it is important to keep magnets clear of watches and electronic
equipment that is sensitive to magnetic fields. Floppy disks and magnetic
tapes are particularly vulnerable, and high-energy magnets can seriously
distort the image on computer terminals and monitors. An engineer who
works with permanent magnets or PM motors will quickly learn to keep
magnet samples in a steel cabinet well away from his credit cards and
computer disks!

M agnet-. Wooden push-rod-

Steel guide

Rotor

Ilg. lfi.5 Method for inserting magnets using slotted Weeper

1 This method waa invented by D.W. Jones of CE. Once inside the steel guide, even
the strongest magnets become completely docile and can be pushed around at willl
16-11
D e s ig n o f b r u sh l e s s p e r m a n e n t - m a g n e t m o t o r s

Permanent magnets require strict adherence to safety procedures at all stages of


handling and assembly
1. Always wear safety glasses when handling magnets. This is particularly important
when assembling magnets into a motor. Whi:n a large pole magnet is being
assembled from smaller magnets, the magnets have a tendency to flip and
jumpunexpectedly, and may fly a considerable distance.
2. Work behind a Plexiglass screen when experimenting with or assembling magnets.
Watch out for trapped fingers, especially with large magnets or high-energy magnets.
3. Avoid chipping by impact with hard materials, tools, or other magnets.
4. Never diygrind rare-earth magnetsthe powder is combustible. In case of fire, use
LP argon or nitrogen dry chemical fire extinguishers: never use water or halogens.
5. PM motors generate voltage when the shaft is rotated, even when disconnected.
This may be obvious to the engineer, but is a potential safety hazard for electricians,
maintenance personnel, and casual observers who may rotate the shaft by hand out of
curiosity. Use suitable warning labels, especially on large machines. Alternatively,
connect a bolted short-circuit across the phase terminals when the motor is
disconnected from the supply and the shaft is decoupled from the load (or prime
mover, in the case of PM generators).
fi. When assembling a PM rotor into a stator, the rotor must be firmly guided and
restrained and the stator must be firmly located. When the rotor is partially entered
in to the stator a very large magnetic force tends to draw it all the way in. This force is
uncontrollable and very dangerous except in the tiniest machines.
7. When transporting or shipping magnets or magnetized rotors, use keepers to
ensure that no flux leaks outside the package.
T able 16.5 Safety w ith perm anent magnets
Magnets are usually held in place by bonding (for example with Loctite
Multibond) or compression clips. In motors with magnets on the rotor,
adhesive bonding is adequate for low peripheral speeds and moderate
temperatures, but for high speeds a Kevlar banding or stainless steel
retaining shell can be used. In motors it is not advisable to make the
magnet an integral part of the structure. Mechanically, the magnet
should be regarded as a passenger for which space and fixturing must
be provided. The important requirements are that the magnet should not
move and that it should be protected from excessive temperatures.
A very wide range of magnet shapes is available and many more can be
designed and produced on a custom basis, but in motors the most

16-12
16. M a g n e t ic m a t er ia l s
common are arcs and sometimes rectangles. Tolerances in the
magnetized direction can be held very close, ^/-O.lmm even for standard
magnets. But if the design permits a relaxation of the required tolerance,
particularly in the dimensions perpendicular to the magnetic axis, this
should be exploited because it reduces the cost of the finished magnets.
Thermal expansion of magnets is usually different in the directions
parallel and perpendicular to the magnetic axis. The coefficients in Table
16.1 are for the direction of magnetization, i.e. along the magnetic axis.
Most magnets have a high compressive strength but should never be used
in tension or bending.
16.2.7 The latest trends in magnet technology
There is still much research into new methods of producing Nd-Fe-B
materials especially with regard to corrosion resistance. There is also
work on new materials such as Br-Fe-B (Br = 12.8kG, Hci = 11.5kOe,
(BH)m*x = 38MGOe) and Sm-Fe-N (r = 19kG, //d = 9kOe, (BH)max =
20MGOe).
There has been a 10-14% growth rate per year for bonded materials over
the last 5 years. A 180C operating temperature bonding material is also
now available. Mitsubishi have developed the HDDR process for making
anisotropic Nd-Fe-B materials (this was developed at the end of 1993).
This is believed by many to be the most important development for
bonded magnets in the last 10 years. The HDDR process makes it
possible to manufacture 20MGOe bonded magnets. There is also a cross-
licensing agreement for the HDDR process with Sumitomo and General
Motors. Activity in magnet research is well reported in IEEE and
specialist conference proceedings.

16.3 Soft magnetic iron


Electrical steels (core steels) used in motors are called "soft" because they
have narrow hysteresis loops, low coercivity, and high permeabilityquite
different from permanent magnets. The high permeability is needed
because the essential function of core steel is to act as a flux guide, and
it should absorb the minimum MMF so that the precious MMF of the
magnet can be focussed where it is needed most, in the airgap.

16-13
DESIGN OF BRUSHLESS PERMANENl'-MACNET MOTORS

Fig. 16.6 Typical DC magnetization curves for electrical steels.

Low coercivity (a narrow hysteresis loop) is required to minimize


hysteresis losses: generally in brushless motors the armature core
experiences an alternating flux as well as high-frequency flux variation
due to PWM. The remanent flux-density of electrical steels is quite high,
but this is because the material has a high saturation flux-density,
necessary to carry' as much flux as possible with the minimum cross
section (and weight). In high-frequency applications the eddy-currents
are minimized by using thinner laminations and high-resistivity steels
(usually silicon steel with 1-3% Silicon content).
There have been great improvements in the quality of electrical steels
over the last 20 years. This has been made possible by improved
manufacturing techniques and a better understanding of factors which
control magnetic properties. Due to the radial magnetization direction
required in the brushless motor, non-oriented steel must be used. There
are many different grades and gages of non-oriented steel available, and
the choice of material is largely a compromise between performance and
16-14
16. M a c n e t ic m a teria ls

cost. Table 16.2 lists the main properties of some of the more common
materials used in brushless motors in the USA. Table 16.3 presents
equivalent information for materials used in Europe.
16.3.1 The DC magnetization curve
In order to calculate the airgap flux and particularly the MMF
requirement of an electric motor it is necessary to know the dc
magnetization curve (or initial magnetization curve) of the soft magnetic
material used in the magnetic circuit. The dc magnetization curve is in
effect the average value of B versus H of the major hysteresis loop. It is
noted that typical elcctrical steels have very narrow hysteresis loops
compared with permanent magnet materials. Fig. 16.6 compares the dc
magnetization curves of a number of common electrical steels.
Grade Saturation Core Low (W/lb)
Induction 15kG and 60Hz (lW/lig - 0.454 W/lb)
(Cauu) 0.014" 0.0185" 0.025"
29 Gauge 2fi Gauge 24 Gauge
M-I5 19,900 1.45 1.68 NA
M-19 19,900 1.58 1.74 2.08
M-36 20,200 1.90 2.05 2.40
M-43 20,400 2.00 2.30 2.70
T able 16.2: L am ination materials (USA)

Grade Saturation Core Loss (W/kg)


Flux Density 1.5T and 50Hz. (lW/kg = 0.45-1 W/lb)
H=5kA/m, 50Hz
(Tcjla) 0.35mm 0.50mm 0.65mm
Transit 330 1.66 2.90 3.15 NA
Losil 600 1.70 NA 5.10 5.50
Newcor 800 1.72 NA 6.20 7.40
T arle 16.3: L am ination materials (E u r o pe )
16-15
Specific Total Loss [W/kg] D e s ig n o f b r u s h l e s s p e r m a n e n t -m a c n e t m o t o r s

Fig. 16.7 Typical core loss data (Transit 355-50 )

16,3.2 Core losses


The soft iron of the stator suffers eddy current and hysteresis core losses.
This is mainly due to the rotating magnetic field of the magnets, but
there will also be a load dependent component of loss due to
commutation and PWM chopping. To minimise these losses it is usual to
construct the stator from laminated electrical steels. The rotor of a
brushless motor can be made of any economical steel such as lead free
machining steel. However, it is usually more practical to laminate the
rotor using the material from the hole of the stator stamping. The
lamination will also reduce eddy current losses in the rotor caused by the
lower PWM chopping frequencies from the inverter.
Traditionally the core loss data provided by manufacturers of lamination
materials is limited to 50Hz or 60Hz at either 10 or 15kG flux density.
16-16
16. M a g n e t ic m a ter ia ls

This data is applicable to AC induction motors but not strictly applicable


to brushless motors. Many steel companies can now provide core loss
data as a function of flux density and frequency. Fig. 16.7 shows the
typical form in which core loss data is presented, i.e. the variation in loss
per unit weight with and frequency. If core loss data is only available
at 60Hz, then this is only usable for 4-pole brushless 3 phase motors
operating up to 1800 rev/min, assuming a sinusoidal airgap flux density
profile. For accurate estimates of core losses with non-sinusoidal flux
densities, higher pole numbers and higher speeds, higher frequency core
loss data is required.
16.3.2 Calculation of coefficients for use in core-loss formulas
In Chapter 9 the modified Steinmetz equation for calculating core losses
was introduced. The following procedure should be followed in order to
extract the required core loss coefficients from manufacturers data:
Equation (9.1) must be divided b y /:
= + Ct fE>l W/lb (16.4)
The data should then be used to plot graphs of P//vs. /fo r three values
of e.g. 1, 1.5 and 2T with / from 50 to the highest frequency of
interest.2 The graphs should be straight lines and can be represented by:
= D +B f (16.5)

The intercept D on the vertical (P /f) axis is given by:


D - Q B e*b4 (16.6)

The intercepts Z)j, Z)2 and for the three values of are substituted
into the logarithm of Equation (16.6), giving three simultaneous linear
algebraic equations for C^, a and b of the form:

The most convenient data format is tables or graphs of total core loss vereus frequency
for different fixed values of Bp.

16-17
D e s ig n o f b r u sh l e s s p e r m a n e n t -m a g n e t m o t o r s

logZ>, = log Q + (<s+ W p lo g f^ i (16-7)

These are solved for log a and tr, is then obtained from log
Three values of C are obtained from the gradients of the three graphs
of P /fy s. / (Equation (16.5)). The average or the highest value can be
taken for Cc. Finally:

C is approximately inversely proportional to t2, where i is the lamination


thickness. This can be used to modify Ce or <7cl for different thicknesses:
for example, if the lamination thickness is doubled Cc and should be
doubled. The loss curves should always be re-plotted from the formula
as a check, and any extrapolation to higher Bp or / should be checked
carefully.
16.3.4 Work Hardening
The punching or stamping or blanking of the laminations from coils of
electrical steel causes work-hardening at the shearing edge and this can
substantially degrade the magnetic properties. To restore the desired
magnetic properties, annealing after punching is required. Essentially,
the annealing takes place in a wet forming gas or hydrogen atmosphere
in which the laminations are soaked for at least an hour at a temperature
as high as 950 C (1750F) and allowed to cool slowly. There can be
problems with this process in that the high temperatures burn most
insulation materials away. Thus most lamination manufacturers control
the annealing atmosphere in such a way that an iron-oxide core plate is
produced on the surface of the laminations which serves as an
interlaminar insulator.
If the shape of the lamination is quite simple as opposed to the multi-
teeth poles of a stepping motor lamination, the use of fully processed
electrical steels can be used. That is, the material is annealed before
punching. This process is particularly suited to the rotor core
laminations.
It is interesting to note that laminations annealed after punching are
widely used in the US for most medium to high performance motors. In
16-18
16. M a g n e t ic m a t er ia l s

Europe and Japan this is not the case and typically fully processed
materials are used with without annealing after punching. The reason for
this is that most American motor companies (at least the small ones) do
not punch their own laminations but buy them from large stamping
houses who specialize in punching electrical steel motor and transformer
laminations. In Europe and Japan many motor companies have their own
punching presses and make their own laminations. Under certain
circumstances this may suit just-in-time delivery practices, but the motors
must be designed such that they can handle the increased losses.
16.3.5 Special steels
Special cobalt-iron alloys such as Vanadium Permendur offer saturation
capabilities of up to 2.4T (24kG). This permits a smaller cross section of
iron in the stator and therefore a smaller and lighter machine. However,
these materials are very expensive and are only used in applications that
require high power to weight ratios, e.g. aerospace motors and
generators. To obtain the high powers, high speeds are also used, thus
the steel is usually used in very thin gauges, sometimes with
interlamination insulators to reduce core losses.

16.4 Measurement of material characteristics


With the increasing sophistication of the design techniques employed in
motor design, there is a need for accurate data to characterise the
magnets and electrical steel. In the case of the magnet material, this will
take the form of the major B/H characteristic covering the 1st, 2nd and
3rd quadrants, at ambient, minimum, and maximum working
temperatures, but other data may also be required including the minor,
recoil, and magnetization characteristics. In the case of electrical steel,
the dc magnetization curve and loss data at different frequencies is
required.
Specialist data and measurements are often made by permanent-magnet
research and development bodies; for example, in the UK, the Magnet
Centre at Sunderland University, and in the USA, the University of
Dayton, Ohio. The entire B /H loop of a permanent magnet is usually
symmetrical and can be measured using special instruments such as the
Hysteresisgraph. The Hysterestisgraph can also be used to measure the

16-19
D e s ig n o f b r u s h l e s s p e r m a n e n t -m a g n e t m o t o r s

full hysteresis loop and more importandy the dc magnetization curve of


both ring and bar samples of soft magnetic materials. Specialist
equipment is also used for measuring iron losses under both sinusoidal
and non-sinusoidal field excitation.

16.5 Copper wire


Not much need be said about the conductors used in brushless motors
as copper is the only practical choice in the vast majority of cases. There
are however some interesting forms of copper conductors such as Litz
wire which Ls a multiple-strand conductor used where eddy-currents
and/or circulating currents in the windings would otherwise be
troublesome. For some very high temperature aerospace motors, hollow
conductor copper wire is used so that a cooling fluid can be pumped
down the center. More details on Litz wire and hollow conductors can
be found in reference.
Copper wire has a resistivity of 0.017241 |i2-m at 20 C, a resistance
temperature coefficient of 0.00393 per C and a density of 8900 kg/m .
Copper wire used in electrical machines is usually classified according to
the insulation material used. Table 16.4 gives a list of the insulation
classification codes together with their maximum permitted operating
temperatures.
Insulation class Y A E B F II C
Maximum permitted 90 105 120 ISO 1S5 180 >180
temperature (C)
T able 16.4: C o ppe r wire insulation classifications

"Magnet wire" is normally supplied in discretegauge sizes, although large


and specialty manufacturers often use "half gauge" and even "quarter
gauge" wire. The three main wire gauge systems are:
1. American Wire Gauge (AWG)
2. Standard Wire Gauge (SWG)
3. Metric Wire Gauge (MWG)
16-20
16. M a g n e t ic m a ter ia ls

Copper is not the only possible material for windings, but it is


overwhelmingly the most popular because of its high conductivity and
ductility, which helps in winding. However, its aging properties and
mechanical strength are not particularly good, especially at high
temperatures, and other materials such as Aluminum, Silver, and
Beryllium Copper are sometimes experimented with.

Reference
1. Parker RJ [1990] Advances in Permanent Magnetism, John Wiley &
Sons, N.Y., ISBN (M71-82293-0.

16-21
Index
AC 1.1-1.4, 1.12, 1.24, 2.2, 2.6, 2.7, 2.20, 2.38-2.40, 2.46, 2.47, 2.49, S.2, 3.3. 3.5, S.12,
3.14, S.29-3.31, 3.33, 3.35, 4.18, 5.9, 5.18, 5.31, 5.63, 6.1-6.4, 6.9, 6.12,
6.13, 6.29, 6.40, 6.47, 7.1, 7.7, 7.8, 7.13, 7.15, 9.1, 11.6, 11.14,11.16,
12.2, 12.3, 13.1-13,3, 13.6. 13.7, 13.11. 13.14, 15.9, 16.14
AC induction motor 1.1
see induction motor
AC servomotor 1.2
Accuracy 1.21, 1.25, 2.47, 3.24, 4.26, 10.32, 12.14, 14.13, 14.18
Adaptive control 14.1, 14.31, 14.33, 14.43, 14.45
Air cooling 12.5
Airflow 15.7, 15.9
Air-over 12.5, 13.14, 15.7
Airgap 1.1,1.19, 2-5, 2.6, 2.11,2.12, 3.1, 3.8, 3.19, 3.20, 3.21,3.23, 3.24.3.25. 3.26,3.51,
3.71, 4.2, 4.5, 4.7, 4.8, 4.10, 4.12-4.16, 4.23, 4.25, 4.28, 4.30, 4.32, 5.1,
5.12, 5.19, 5.21, 5.35, 5.42-5.55, 5.61,6.1, 6.3-65, 6.12, 6.13, 6.17, 6.24,
6.25,6.28-6.33, 6.37, 6.46, 7.3. 7.12, 8.4-8.10, 8.12, 8.13,8.16, 9.2,10.2,
12.2-12.4,12.6,12.12,12.15-12.17,12.19-12.23,12.25-12.28, 13.4,13.6,
13.9, 13.-11, 13.14, 13.19-13.22, 16.4, 16.7, 16,8, 16.11, 16.12, 16.14
Airgap flux 2.5, 3.24, 4.2, 4.10,4.14, 4.16. 4.30, 4.32, 5.1, 5.12, 5.21, 5.35, 5.43. 6.4, 6.5,
6.12, 6.25, 6.28, 6.3M.33, 6.46, 7.12, 8.&4.10, 8.13, 9.2, 12.12,
12.19-12.21, 12.25-12.28, 16.12, 16.14
Airgap length 3.24, 4.2, 4.13-4.15, 4.25, 4.28. 5.61, 6.3, 13.9, 16.8
Airgap shear stress 12.2, 12.4, 12.6
Alger 5.58, 5,66
Aliasing 14.35
Alnico 2.19, 3.23, 4.5, 16.7, 16.8
Ampere 1.1-1.3, 1.17, 2.15, 2.20, 2.46. 3.4, 3.6, 3.26, 4.9, 4.19, 4.20, 4.22, 5.18-5.21,
5.25, 5.28, 5.31, 5.36, 5.56, 6.4-6.7. 6.1M.13, 6.16, 6.20, 6.23, 6.24,
6.28, 6.31,6.35,6.39,6.42-6.44, 7.1, 7.14,7.15,11.2,12.21-12.23,12.26,
12.27, 16.7, 16.8
Ampere-conductors 3.6, 4.19, 4.20, 4.22, 5.18, 5.19, 5.21, 5.25, 5.31, 6.4, 6.6, 6.7, 6.10.
6.16, 6.20, 6.28, 12.22, 12.23
Amplifier 2.43, 2.44, 6.43, 14.11, 14.12, 14.22
Analog converter 14.34, 14.41
Annealing 13.22-13.24, 16.15, 16.16
Anti-windup 14,32, 14,34
Applications 1;1, 1.3, 1.23, 1.25, 2.1, 2.5-2 6, 2.9-2.11, 2.13, 2.18, 2 20. 2.39, 2.43, 2.45,
2.47, 2.49, 3.1-3.3, 3.5, 3.9, 3.25-3.28. 3.33, 3.71, 3.76, 4.33, 5.18, 5.32,
6.47, 6.48, 8.3, 8.19, 8.20, 9.9, 10.35, 11.1, 12.5, 12.14, 12.18, 13.1,
13.19, 14.1, 14.9, 14.13, 14.14, 14.19, 14.21, 14.46, 15,2, 15.14, 15.22,
16.4, 16.6-16.8, 16.11, 16.16
Arc 2.3, 3.8, 3.12, 3.19-3.21. 3.25, 3.33, 3.45, 3.47, 3.50, 3.51, 4.12, 4.24, 4.25, 5.1, 5.5,
5.8, 5.14, 5.19-5.21, 5.24-5.26, 5.28-5.32, 5.35,5.60, 5.65, 6.3, 6.29,7.12,
8.1, 8.2, 8.5-8.7, 8.11, 8.12, 8.15, 9.5, 9.6, 10.15, 13,4, 13
Arc magnets 3.25
Armature 1.4, 2.8, 3.3. 5.20, 3.23, 3.35, 4.12,4.18, 5.39, 5.40, 5.46, 5.55, 5.63, 6.3, 6.12,
6.13, 6.23, 6.24, 6.28, 6.90-6.34, 6.38, 6.40, 7.S-7.5, 8.20, 11.2, 11.9,
12.20-12.28, 14.5, 15.9, 16.2. 16.8, 16.11
Armature reaction 3.20, 4.18, 5,39,5.40, 5.46, 5.55, 5.63, 6.3, 6.12, 6.13, 6.23-6.24, 6.28,
6.30-6.34, 6,40. 7.4, 7.5, 11.2, 11.9, 12.20-12.22, 12.24-12.28, 16.2
Armature-reaction flux 5.55, 6.23, 6.24, 6.33, 6.34
Augmentation of stator teeth 9.6, 9.7
Auto-tuning 14.33
Axial gap 1.4
Axially laminated motor 18.1 IT
Axis 1.7, 2.46, 3.21, S.42, 3.43, 3.45, 4.6-4.8, 4.23, 5.1-5.3, 5.5, 5.35, 5.47. 5.63. 6.2, 6.3,
6.8-6.10. 6.17, 6.18,6.20,6.22-6.25, 6.2M .38,6.40,6.42-6.44.6.46, 7.15,
8.5, 8.9,8.16, 11.4,12.21, 12.25-12.27, 13.19, 14.16, 14.23-14.25.14.32,
16.10,16.11, 16.14
Back-EMF1.2,1.7-1.9,1.13,1.14,1.23, 2.14, 2.15, 2,17, 2.22, 2.23, 2.30, 2.33, 2.34, 2.43,
2.44, 2.46-2.48, 3.6, 3.15, 3.16, 3.18, 3.19, 3.33, 3.38, 3.45-3.47,
3.49-3.51, 3.71, 3.74, 4.18, 4.28, 5.3, 5.5, 5.8, 5.9, 5.12, 5.21, 5.25, 5.35,
5.36, 5.40-5.42. 5.63, 5.65, 6.2, 6.4, 6.36, 6.42, 6.46, 7.1-7.5, 7.7, 8.1-8.3,
8.11-8.13, 8.20, 9.2, 9.9, 10.1, 10.28-10.32, 11.2,11.3,11.9, 12.9,12.12,
12.20, 13.8, 13.9, 13.11, 13.13, 13.19, 13.25, 14.7
Back-EMF constant 3.18, 3.19, 5.12, 5.42, 7.1, 7.3, 8.1, 10.1, 10.30, 13.19
Back-iron 2.11, 3.6, 3.25, 3.34, 4.2
Bandwidth 1.19, 2.22, 2.47, 3.2, 11.12, 14.12, 14.15, 14.18, 14.26, 14.32, 14.35
Bare wire 13.13, 13.26
Base interval 2.31-2.33, 10.1, 10.2, 10.4, 10.13, 10.20, 10.22, 10.24
Base speed 1.12
Bearings 1;17,1.20, 1.22, 2.2, 2.6, 2.8, 3.8,3.17,4.33, 5.39, 11.6,13.2,13.15,13.25.15.1,
15.2
Bekey-Robinson 2.5
Bifilar-wound 3.4
Bifurcated teeth 4.27, 4.28, 8.14
B1L 3.16
BJT 2.20
Blocking 2.18, 2.19, 3.35
Black body 15.5
BLV 5.19, 5.36, 8.3^.5, 8.9-S.ll
Bode diagram 14.16, 14.17, 14.21, 14.26, 14.27
Bonded ferrite 2.3
Bonded ring magnet 2.3, 2.8, 3.8
Bonding 3.28, 3.29, 16.10, 16.11
Boules 5.64, 5.66, 8.3, 8.19, 8.20
Boundaryelemcnt method 5.38, 5.42, 5.61, 5.66, 8.2
Brake 11.6, 11.9
Braking 2.17, 2.44, 2.45, 3.2, 6.47, 14.1, 14.12
Brushes 1.1
C 14.34
Carbon steel 2.2, 3.25, 13.21, 15.13
Carter coefficient 3.21, 4.13, 5.44, 8.12
Centrifugal blower 2.9

xviii
I n d ex
Centrifugal loading 4.28
Centrifugal pump 1.12
Centrifuges 1.17
Ceramic 4.5, 15.13, 16.8
Ceramic magnets 4.5
Chamfer 3.8, 4.28
Characteristic equation 14.22
Charge 2,17
Chemicals S.2, 3,32, 16,10
Chopping 1.17, 1.19, 2.15, 2.21, 2.23, 2.24, 2.26, 2.27, 2.30, 2.32-2.35, 2.37, 2.41, 2.43,
3.25,4,18,5.32, 5.39,5.41,7.5.8.1,10.7-10.14,10.16-10.18,10.22-10.24,
10.26-10.28, 16.13
Chording 3.36, 5.29, 5.30, 6.3, 6.14, 6.16, 6.18, 6.19
Circle diagram 17,11 et waj.
Circuit 1.10, 1.14, 1.19, 1.24, 1.25, 2.1, 2.9-2.11, 2.15-2.17, 2.22-2.28, 2.30, 2.31, 2.41,
2.45, 2.47, 2.48, 3.4,3.14, 3.15, 3.17, 3.20,3.21,3.23, 3.29,3.S0,4.S4.9,
4.11-1.16, 4.19, 4.254.27, 4.32, 5.14, 5.21. 5.32, 5.33, 5.38, 5.40, 5.41,
5.43, 5.58, 5.61, 6.3,6.12, 6.22, 6.24, 6.25, 6.3W.35, 6.45.6.47, 7.4, 7.8,
8.2. 8.4, 9.2, 10.2, 10.3, 10.7, 10.8, 10.10, 10.12-10.15. 10.20, 10.21.
10.23, 10.24, 10.27, 10.28, 11.2, 11.5, 11.9, 12.21, 12.22, 12.26, 13.3,
13.4, 13.16, 14.1, 14.7, 14.9, 14.11, 14.14, 14.18, 14.34, 14.41,
15.10-15.12, 15.15, 15.21-15.23, 16.1, 16.4, 16.8, 16.10, 16.12
Cleat 3.29. 13.25
Closed-loop control 3.2, 14.10, 14.15, 14.20
Closed-loop transfer function 14.25, 14.26
Cobalt 3.8, 3.28, 4.3, 4.9, 6.47, 6.48, 15.13, 16.4, 16.16
Cobali-Sainarium 4.3, 4.9
Coefficient2.3,3.17,3.19-3.21,3.23,3.24,3.27,4.8-4.11,4.13-4.15,4.24,4.25, 5.1S, 5.44,
5.55-5.59, 6.9, 6.34, 8.12, 12.2, 13.4, 15.2, 15.5-15.8, 16.7, 16.8, 16.17
Coenergy 3.17, 4.27, 5.36, 5.37
Coercive force 16.8
Coercivity 3.23, 3.25, 3.27, 4.6, 4.8, 4.9, 4.15, 4.16, 4.18, 5.64, 6.12, 12.22, 13.4, 16.2,
16.3, 16.7, 16.8, 16.11
Cogging 1.19, 2.1, 2.3, 2.11, 2.13, 2.47, 3.3, 3.7, 3.8, 3.10, 3.12, 3.13, 3.24, 3.34, 3.51,
3.74.4.26-4.28,4.33,5.57,5.62,6.21,8.10,8.19,12.14,13.7,13.9,13.18,
13.22, 13.23
Coil 1.7,1.9,1.10,1.14, 2.1,2.6, 2.10-2.12, 3.1, 3.5, 3.10,3.12, 3.14-3.15,3.20, 3.23,3.28,
3.30, 3.31-3.36, 3.39-3.51, 3.70, 3.71, 4.1, 4.5, 4.10, 4.20, 4.21, 5.2, 5.3,
5.5, 5.6, 5.8, 5.9, 5.11-5.14, 5.20, 5.21, 5.29, 5.30, 5.35, 5.36, 5.43-5.49,
5.51-5.55, 5.59, 5.60, 5.64, 5.65, 6.1, 6.3, 6.5-fi.8, 6.13-6.21, 6.25. 7.11,
7.12, 8.1, 8.2, 8.4, 8.5, 8.9-8.11, 8.17, 9.7,10.15,11.12, 12.9,13.1, 15.2,
13.6, 13.8, 13.11-15.13, 13.15, 13.16, 13.22, 13.23,13.25, 13.26, 14.18,
15.4, 15.5, 15.24, 16.8, 16.15
Coil pitch 5.3. 5.9, 5.20, 5.29, 8.1, 8.4, 10,15
Coil span 3.14, 3.35. 3.36, 3.39, 3.41, 3.42, 3.49, 3.51, 5.20, 6.16, 8,2
Coil winding 5.9
Coils per phase 3.36, 3.42, 3.44, 5.12, 5.60, 13.8, 13.13, 13.26

xix
Coils per pole 3.36, 3.39, 3.49, 6.13, 6.17, 6.25, 13.8
Combinatorial logic 2.45
Commutation 1.2, 1.5,1.12,1.18-1.23, 2.15, 2.19, 2.20, 2.32-2.35. 2.42, 2.43, 2.45, 2.47,
3.1, 3.4, 3.6.3.32, 3.50,3.51, 3.71,5.5, 5.15,5.17, 5.19,5.31, 5,32. 6.41,
7.2, 7.6, 8.1,10.1,10.2, 10.4,10.7-10.9, 10.13-10.17,10.20,10.23,10.24,
10.28, 10.30,10.31, 10.33,11.1,11.2, 11.6.11.7, 11.9,11.11,13.6,13.7,
13.11, 13.16, 13.25, 16.13
Comparator 2.22
Compensator 14.11, 14.12, 14.25-14.27
Complex frequency 14.2
Compressors 1.12, 15.1, 15.2
Computer disc 3.71, 3.73, 3.74, 13.16, 13,18
Concencrated winding 6.18, 7.14
Concentric winding 3.39, 3.40, 5.9, 6.17, 13.8
Conduction 1.13,1.19,2.13, 2.17-2.19.2.24,2.25. 2.32,2.33, 2.35,2.38, 2.41,2.44, 3.32,
4.22, 5.5, 5.17-5.19, 5.24. 5.25, 5.41, 7.2, 10.9, 10.16, 10.24, 10.27,
10.28, 10.30, 11.9, 15.1, 15.2, 15.4, 15.5, 15.8-15.11
Conductivity 3.15, 15.3-15.6, 15.10, 15.11, 15.23, 15.24, 16.17
Conductor 1.1-1.4,1.7,1.17, 2.12,2.13, 2.15,2.46,3.1,3 4-3.6,3.16-3.18, 3.21,3.26,3.27,
3.30,3.32, 3.33,3.38,3.70,4.2,4.19,4.20,4.22, 5.5, 5.9, 5.18-5-21,5.25,
5.28, 5.31, 5.35, 5.36, 5.39, 5.43, 5.44, 5.46, 5.48, 5.49, 5.51, 5.53-5.65,
6;1, 6.M.7, 6.10-6.13, 6.15, 6.16, 6.20. 6.23-6.25, 6.28, 6.31, 6.35, 6.44,
7.11, 7.12, 7.14, 7.15, 8;4, 8.9, 8.10, 12.6, 12.15, 12.22, 12.23, 12.26,
12.27, 13.11, 13.13, 13.15, 13.25, 15.1, 15.4, 15;5, 15.11, 15.12, 15.23,
16.8, 16.16
Consequent-pole 2.3, 3.24, 3.36, 4.21, 5.11-5.13, 13.8
Constant power operation 1.12, 2.3, 6.4, 6.46, 17.1 et seq.
Constant-power speed range
see CPSR
Contact resistance 15.1, 15.4, 15.5
Continuous 1.16, 1.17, 2.10, 3.2. 3.21, 3.31, 3.76, 5.18, 5.31, 5.32, 6.42, 6.47, 7.2,11.7,
13.13, 13.14, 14.34, 14.36-14.38, 14.40, 15.15, 15.17, 16
Control 1.1, 1.2, 1.5, 1.11, 1.12, 1.18, 1.23, 1.25, 2.9, 2.11, 2.14, 2.15, 2.19, 2.23, 2.24,
2.30-2.36, 2.38-2.47, 2.49, 3.2, 3.4, 3.5, 3.25, 3.31, 3.76, 5.18, 6.2, 6.3,
6.11, 6.12, 6.43, 6.48, 7.1, 7.3, 8.1, 10.10, 10.28, 11.6, 11.7, 11.9, 12.7,
12.10, 13.16, 14.1, 14.7, 14.9-14.16, 14.18, 14.20, 14.22, 14.23, 14.25,
14.26, 14.28-14.36, 14.38-14.43, 14.45, 14.46, 15.1, 15.14, 16.11, 16.15
Convection 12.5, 15.1, 15.2, 15.6-15.12
Converter 1.19, 1.25, 2.43, 6.38, 6.43, 10.16, 10.18, 10.28, 14.34, 14.41
Conveyors 1.12
Cooling 1.17, 2.2, 2.13, 3.3, 3.33, 4.32, 11.1, 11.8, 12.4-12.6, 13.1, 13.14, 15.1-15.2,
15.5-15.9, 15.11, 15.12, 16,7, 16.16
Copper 1.17,1.20, 2.12, 2.47, 3.5, 3.15-3.17, 3.28. 3.32, 3.70, 4.31,5.11, 5.12, 5.21, 5.31,
5.39, 5.61, 5.63, 6.3, 9.1. 13.13, 13.14, 13.17, 13.26, 15.1, 15.4, 15.5,
15.8, 15.10, 15.12-15.14, 15.23, 15.24, 16.8, 16.16, 16.17
Copper losses 2.12, 3.32, 5.31, 5.39, 9.1, 13.14, 13.26, 15.10
Core 1.17, 2.1, 2.1S, 3.25, 3.28, 3.33, 3.34, 3.71, 3.73, 4,1. 4.2, 4.33, 5.60, 8.3. 8.4, 8.10,
8.20, 9.1-9.3, 9.6-9.9,11.4,12.15,12.20, 12.21,12.27, 13.3, 1S.7,13.14,
13.15, 13.23, 13.24, 15.4, 15,5, 15.10, 16.11-16.16

XX
Index
Core losses 1.17, 2.1, 3.28, 3.35, 3.34, 4.2, 4.33, B.3-8.4, 8.10, 8.20, 9.1, 9.2, 9.7. 9.9,
12.20, 12.21, 13.3, 13.7, 13.14, 13.15, 13.23, 13.24, 15.10, 16.12-16.14,
16.16
Core plate 3.28, 3.33, 16.15
Core steel 15.15, 15.5, 16.11
CPSR 17.2 et seq., 18.7 tt *eq.
Critical damping 14.24
Cross-inagnetizing 6.33, 12.24, 12.25, 12.27
Cuffs 3.29
Current 1.1, 1.2, 1.4, 1.8, 1.9, 1.11-1.16, 1.18, 1.19, 1.22, 2.3, 2.6, 2.11, 2.12, 2.14-2.27,
2.29-2.35, 2.37-2.48, 3.1, 3.4, 3.6, 3.8, 3.15-3.17, 3.19, 3.20, 3.23,
3.25-3.27, 3.32, 3.33, 3,50, 3.71, 4.M.4, 4.7-4.9, 4.12, 4.1&4.20, 4.22,
4.26-4.29, 4.31,4.33, 5,4,5.5, 5.9, 5.12,5.17, 5.18, 5.21, 5.23-5.27,5.29,
5.31-5.34,5.36-5.41,5.43,5.44,5.46,5.47, 5.54,5.61,5.63-5.65,6.1,6.4,
6.8, 6.12, 6.14, 6.16, 6.28, 6.30, 6.32, 6.36^.44, 6.46, 6.47, 7.1-7.4,
7.6-7.10,7.12,7.13,7.15,7.16, 8.1-8.3, 8.11, 8.16, 8.19, 8.20,9.1,9.5-9.7.
10.1,10.2,10.4,10.7-10.11,10.13-10.16,10.18,10.21-10.25,10.27-10.35,
11.1, 11.2, 11.5, 11.6, 11.8, 11.9, 11.12-11.14, 12.2-12.5, 12.12, 12.15,
12.18-12.21, 12.23, 12.25-12.27, 13.4, 13.8, 1S.13-1S.16, 13.19, 13.25,
13.26,14.1,14.5, 14.7,14.14,14.31,14.45,15.2,15.4,15.8-15.10,15.12,
15.17,15.18, 15.24, 16.6, 16.8, 16.13
Current density 3.1, S.32,4.31, 5.61, 5.63,5.64. 6.4,12.2,12.4, 12.5,12.19, 12.23, 13.13,
13.14, 1S.26, 15.4, 15.8, 15.9, 15.12
Current regulator 2.15, 2.38, 2.46
Current transducer 2.22, 11.6
Current-regulated 2.43
Curve 1.12,1.13, 1.15-1.18, 2.47, 3.20, 3.21, 3.23, 3.25-3.27, 3.51, 3.71, 4.16, 4.17, 4.33,
5.35, 11.1, 11.2, 11.5, 11.7,12.17, 12.20,13.4-1S.6, 13.14, 13.15, 13.19,
13.20, 14.45, 15.7, 16.1-16.4, 16.12, 16.16
Cutting 3.28, 5.36, 14.14, 15.24
Cycle 2.15, 2.17, 2.21, 2.22, 2.24, 2.26, 2.27, 2.34, 2.37, 2.43-2.45, 3.18, 4.31, 5.5, 6.43,
9.1, 9.3, 10.11, 10.13, 10.18, 10.22, 10.27, 10.28, 10.32, 10.33, 14.24,
15.14-15.16, 15.18-15.20, 15.22
d-axis 5.1-5.3, 5.35, 6.23-6.25, 6.28-6.30, 6.33-6.38, 6.40, 6.44, 6.46. 12.21, 12.25, 12.27
Damping 10.24, 14.7, 14.9, 14.13, 14.24, 14.25, 14.38, 14.43
dB 8.7,9.2, 9.3, 9.6, 14.16, 14.18, 14.21
DC 1.1-1.12, 1.14-1.20, 1.25, 2.6, 2.8, 2.10, 2.15, 2.19, 2.23, 2.27, 2.30, 2.32, 2.33, 2.35,
2.38-2.41, 2.44, 3.1-3.6, 3.19, 3.26, 3.29, 3.33, 3.35, 3.71-3.76, 4.4, 4.5,
4.18,4.22, 4.24, 4.33,5.1, 5.2,5.6,5.18, 5.25-5.27,5.32, 5.33, 5.36,5.39,
5.40, 5.57, 5.66, 6.12, 6.41, 6.47, 7.1-7.11, 8.1, 8.19, 8.20, 9.1,9.2, 9.9,
10.1, 10.2, 10.13, 10.14, 10.20-10.22. 10.26,10.28, 10.30, 10.34,10.35.
11.1,11.2, 11.4-11.7,12.17, 12.20,12.21, 1S.1,13.3,13.4,13.11,13.14,
13.19, 14.5, 14.7, 14.16-14.18, 14.28, 14.31, 14.41, 14.42, 14.46, 15.2,
15.9, 15.24, 16.11, 16.12, 16.16
DC commutator motor 1.2-1.9, 5.18, 7.3, 7.7
Deadbeat response 14.40
Decibcl 14.16

xxi
Demag current 3.26, 3.27
Demagnetization 1.16, 1.17, 1.18, 2.3, 3.1, 3.23-3.27, 3.74, 4.54.9, 4.12, 4.14-4.16, 4.1&.
4.20, 4.22, 4.27, 4.28, 5.3&-5.40. 5.62, 5.64, 6.33, 16.4, 16.7,6.37-6.38,
6.40,6.44,6.47,11.4,11.5,12.20,12.21,18.25,12.27,12.28.13.4,13.15,
13.16, 13.19, 16.1, 16.2, 16.3, 16.4, 16.7, 16.8
Demand signal 10.28, 14.10, 14.11, 14.14, 14.18, 14.20, 14.21, 14.31
Design 1.2, 1.3, 1.5. 2.1-2.3, 2.5, 2.6, 2.9, 2.11, 2.13, 2.47, 3.1, 3.3. 3.7, 3.8, 3.10, 3.14,
3.17-3.21, 3.23-3.26, 5.28, 3.32, 3.34, 3.36,3.71, 4.1, 4.2, 4.9, 4.15, 4.22,
4.29, 4.30, 5.1, 5.13, 5.33, 5.38, 5.39, 5.42, 5.58, 5.63, 5.66, 6.21, 6.33,
6.42, 6.47, 7.1, 8.1-8.S, 8.20, 9.3, 11.1, 12.1, 12.2, 12.6-12.10, 12.13,
12.14,12.19,12.20,12.28,13.1-13.4,13.6-13.9,13.14,13.16-13.19.13.22,
13.23, 13.25-13.27, 14.1, 14.7, 14.9, 14.10, 14.16, 14.18, 14.23-14.26,
14.30, 14.31, 14.33, 14.34, 14.38-14.40, 14.43, 14.46, 15.1, 15.23, 16.8.
16.10. 16.16
Detent torque 3.34, 13.7, 13.18
Dielcctric strength 3.32. 15.13
Digital control 1.23, 2.19. 14.11, 14.34, 14.38
Digi(a 1-to-analog converter 14.41
Digital PID 14.34, 14.40, 14.41
Dimpled staclu 13.25
Diode 2.16, 2.17, 2.19, 2.21, 2.23, 2.26, 2.27, 2.30, 2.33-2.38, 6.41, 10.7, 10.8, 10.10,
10.15
Direct axis
see d-axis
Direct conductor cooling 12.6
Disc drive 3.73, 3.74. 11.7, 13.1, 13.16, 15.18, 13.19
Dissipation 6.42, 15.15, 15.17-15.20
Distribution factor 3.45, 6.18, 6.19
Dominant poles 14.24
Double-layer winding 3.36, 5.9, 5.21, 5.30, 13.23
Drive 1.2, 1.3, 1.11, 1.12,1.17, 1.20, 1.24, 2.1, 2.5, 2.9, 2.10, 2.19, 2.20, 2.23, 2.26, 2.27,
2.35, 2.37, 2.39, 2.40, 2.42, 2.43, 2.452.47, 2.49, 3.2, S.4-3.6, 3.15, 3.24,
3.26, 3.27, 3.46, 3.47, 3.49, 3.50, 3.71, 3.73. 3.74. 4.2, 4.8, 4.26-4.28,
4.33, 5.32, 7.4, 7.16, 9.1, 10.14, 10.22, 11.1, 11.2, 11.4, 11.5, 11.7.
11.11-11.14, 12.9, 13.1,13.4, 13.8, 13.9, 13.16, 13.18, 13.19. 14.1, 14.7,
14.9, 14.11, 14.17-14.19, 14.32, 15.9, 16.3
Driver 2.19, 13.25, 14.14
Dummy slots 4.27
see bifurcated teeth
Dust 1.23, 3.2, 3.32
Duty-cycle 2.15, 2.21, 2.22, 2.26, 2-27, 2.34, 2.37, 2.43, 2.44, 4.31, 6.43, 10.11, 10.13,
10.18.10.22,10.27,10,28, 10.32,10.33,15.14-15.16,15.1
Dynamic braking 2.17, 2.44, 2.45, 14.1
Dynamics 14.9, 14.12, 14.22. 14.26, 14.30, 15.7
Dynamometer 1.13, 4.33, 7.4, 11.2, 11.6, 11.7, 11,9, 11.12-11.16
ECM 2.6
Eddy current losses 16.13
Eddy-currents 1.4. 1.17. 2.6. 2.11, 2.12, 3.25. 3.33, 4.2, 5.61. 9.1, 9.5-9.7, 11.6, 12.27,
13.14

XXll
I n d ex
EDM 3.28
Effective airgap 3.21. 5.52. 5.61, 6.30, 8.10, 8.12
Efficiency 1.1, 2.6, 2.9, 2.45, 3.28, 5.31, 5.32, 5.39, 8.2,10.1-10.2,11.7,11.12,12.1,12.3,
12.5, 13.15, 14.46, 15.2, 15.14
Electric 1.12, 2.6, 2.12, 3.6, 3.14, 3.15, 3.17, 4.3, 5.36, 5.40, 5.61, 7.1, 10.34, 12.2-12.5,
12.28, 14.14, 14.40, 15.5, 15.10, 15.21, 16.8, 16.12
Electric loading 2.12, 3.6, 5.61, 12.2-12.5
Electric vehicles 1.12, 14.14
Electrical degrees 2.15, 2.17, 2.30, 2.45, 3.4, 3.18, 3.43, 3.48, 3.50, 3.51, 3.71, 4.12, 5.6,
5.11, 5.20, 5.28, 5.32, 5.46, 6.16, 6.21, 6.22, 6.31, 7.6, 7.12, 7.15, 8.13,
10.1, 10.3, 12.26
Electrical radians 3.18, 4.12, 5.65, 6.3, 6.4, 6.6, 6.&6.10, 6.17, 6.20, 8.5. 8.10, 8.12, 9.7
Electrical steel 9.1, 12.10, 16.1, 16.11-16.13, 16.15, 16.16
Electro-discharge 3.28
Electromagnetic 1.1
Electronic 1.20, 1.22, 1.23, 1.25, 2.9, 2.14-2.16, 2.18-2.20. 2.27, 2.39-2.41, 3.15, 3.19,
3.26, 5.1, 10.1, 10.2, 11.13, 11.14, 14.5, 14.18, 16.9
Electronic wattmeter 11.14
Electronically commutated motor 2.6
Ellipse diagram 17.11 et v.q.
Embedded 2.5, 3.3, 3.7, 3.8, 3.24, 15.24
EMC 3.2, 12.1
EMF 1.2.1.7-1.9,1.11,1.13,1.14,1.22,1.23, 2.14,2.15, 2.17, 2.22, 2.23, 2.30, 2.33. 2.34,
2.43, 2.44, 2.46-2.48,3.6, 3.15-3.20, 3.33,3.38,3.45-3.51,3.71, 3.74, 4.1,
4.18, 4.28, 5.1, 5.S-5.9, 5.12-5.21, 5.24-5.26. 5.28-5.30, 5.32, 5.34-5.36,
5.38, 5.40-5.42, 5.63, 5.65, 6.1-6.4, 6.9, 6.12,6.13, 6.16, 6.19, 6.22, 6.36,
6.42, 6.46,7.1-7.5,7.7,7.10-7.13, 7.15, 7.16, 8.1-8.6, 8.9-8.17, 8.19, 8.20,
9.1-9.3, 9.6, 9.7, 9.9, 10,1, 10.2, 10.12, 10.15, 10.22, 10.25-10.33, 11.2,
11.3,11.9,11.12,12-3,12.9,12.10,12.12,12.20,13.8,13.9,13.11,13.13,
13.19, 13.25, 14.5, 14.7
EMF constant 1.2, 1.13, 3.18, 3.19, 5.12, 5.35, 5.38, 5.42, 7.1, 7.3, 8.1, 10.1, 10.30,
13.11, 13.19, 14.5
EMI 2.39
Encapsulation 15.5, 16.6
Encoder 1.19, 1.21, 1.22, 1.24, 1.25, 2.45, 6.11, 11.5, 11.7, 14.10, 14.13, 14.14, 14,34,
14.41
End turns 3.12, 3.32, 3.35, 13.6, 13.22
End-windings 2.12, 5.9, 5.11, 5.60, 6.3, 15.24
Energy product 3.23-3.25, 4.15, 16.1, 16.4, 16.7
Epoxy 2.8, 3.28-3.31, 11.7, 12.6, 13.6, 13.19, 13.23, 13.25, 15.13
Error amplifier 2.43, 14.11, 14.12, 14.22
Error signal 2.41-2.43, 14.12, 14.20, 14.21, 14.31, 14.41
Exterior rotor 1.3, 1.4, 2.1, 2.2, 2.8-2.10, 3.1, 3.3, 3.6, 3.10, 3.25, 3.35, 3.73, 3.74, 5.21,
12.6, 15.2, 13.16, 13.18, 13.26
Faraday 3.15
Fault 2.41, 3.2, 3.36, 10.2, 11.12, 16.8
Feedback 1.1, 1.18-1.20, 1.24, 2.22, 2.30, 2.43, 2.47, 6.11, 11.6, 11.7, 11.9, 13.1,

xxiii
14.10-14.14,14.19,14.20,14.22,14.23,14.26,14.28,14.31,14.32,14.36.
14.46
Fertile 2.3, 2.5, 2.6, 2.8, 2.10, 2.13, 3.3, 3.8, 3.23, 3.24, 3.26, 3.27, 3.74, 4.3, 4.9, 4.16,
11.7, 12.4, 12.9, 12.22, 13.2, 13.4, 13.11, 16.4, 16.6-16.8
Fiberglass 3.35
Field-oriented control 1.12, 2.46, 6.2, 6.11
Field-weakening 6.46, 17.1 ft vq., 18.1 ft vq.
Finite element method 8.17, 8.20, 12.13, 12.14, 12.27, 12.28
Finite maximum speed 17.11
Finned 1.17, 15.9
Firmware 14.34
First-order hold 14.35
Floppy disc 2.1, 13.16
Flux 1.7-1.9, 1.11, 1.13, 1.17, 1.22, 2.5, 2.6, 2.11-2.14, 2.46, 3.1, 3.3, 3.7, 3.8, 3.15-3.21,
3.25-3.26,3.33, 3.34, 3.71,3.74,4.1-4.17,4.19,4.23-1.33,5.1-5.4,5.6-5.9,
5.12, 5.13, 5.19, 5.21, 5.28, 5.29, 5.31, 5.32, 5.35-5.37, 5.39, 5.42-5.56,
5.61-5.63,5.65,5.66, 6.1-6.10, 6.12-6.17, 6.19, 6.22-6.25,6.28-6.34, 6.37,
6.38, 6.40, 6.42, 6.46, 7.11-7.15, 8.2-8.11, 8.13-8.20, 9.1-9.3, 9.5-9 7,
10.28.11.2, 11.9,11.12, 12.3-12.5,12.12,12.13,12.18-12.28,13,2, 13.4,
13.6,13.9-13.11,15.14, 13.19-13.23, 13.25,13.26,15.9, 16.1,16.2,16.4,
16.7, 16.8, 16.10-16.14
Flux concentration 2.5
Flux distribution 5.1, 5.4, 5.19, 5.21, 5.28. 5.29, 5.35, 5.46, 5.47, 5.49, 5.50, 5.54, 5.62,
6.2, 6.13-6.16, 6.22, 6.24, 6.25, 6.31. 7.15, 8.5, 8.7, 9.2. 12.
12.20-12.23, 12.25-12.27, 13.11
Flux linked 3.15, 5.8, 6.6
Flux-linkage 1.7-1.9, 1.11, 1.13, 3.15, 3.15, 3.17, 3.18, 5.3, 5.4, 5.6-5.9, 5.35-5.37, 5,43,
5.44, 5.46, 5.47, 5.51, 5.52, 5.54-5.56, 5.65, 5.66, 6.1, 6.2, 6.4, 6.6.9,
6.16, 6.17, 6.22, 6.23, 6.31, 6.32, 8.4, 8.9, 12.19, 12.20
Flux-weakening 17.2 ft seq.
Fly winder 13.18
Forced cooling 1.17, 11.8
Four-quadrant 2.17, 14.1
Fracdonal-slot 3.10, 3.12, 5.34, 3.40, 3.41, 3.44, 3.48, 6.17, 6.19, 12.9
Frequency 1.1,1.19,1.21, 2.15, 2.18, 2.20-2.23, 2.34, 2.38,2.43, 2.45, 2.47,3.2, 3.6, 3.33,
3.51, 4.27,5.5,5.9,6.11, 6.42,6.43,6.4M.47, 9.1,9.2,9.5,10.11,10.28,
11.4, 13.6, 13.7, 13.14, 14.1-14.3, 14.6, 14.8, 14.15-14.18, 14.20, 14.21,
14.24-14.26, 14.32, 14.35, 14.3&44.41, 15.9, 16.11, 16.14
Frequency response 14,1, 14.15, 14.16, 14.20
Friction brake 11.6
Full-bridge 2.16, 2.23-2.27, 2.39, 3.4, 5.32
Function 1.10, 1.12, 2.41, 2.46, 4.2, 4.10, 4.27, 4.33, 5.1, 5.3, 5.32, 5.64. 5.65, 6.2, 7.1,
8.5, 8.6, 8.9, 8.12, 8.15, 9,3, 11.8, 12.14, 12.18, 14.2, 14.3, 14.5-14.10,
14.12,14.14,14.16,14.18-14.26,14.28,14.29,14.35,14.36,14.40,14.43,
14.46, 15.4, 15.12, 15.15, 15.19, 16.1, 16.7, 16.11, 16.14
Fundamental 1.5,1.11, 2.1, 2 22, 2.38, 2.46, 3.14, 3.17, 3.33,3.45, 5.1, 5.5.5.8, 5.9, 5.18,
5.24, 5,33. 5.36, 5.61, 5.62, 6.1-6.3, 6.9, 6.10, 6.13, 6.1
6.21-6.24, 6.3^6.32, 6.46, 7.2, 7.13-7.16, 8.9, 9.1, 9.2, 9.5, 12.3, 12.24
Fundamental ampere-conductor distribution 2.46

xxiv
I n d ex
Fundamental winding factor 6.22
Gain 1.12, 1.19, 2.41-2.43, 2.47, 7.1, 10.12, 14.6, 14.9-14.12, 14.16-14.18, 14.20-14.26,
14.28, 14.29, 14.31, 14.32, 14.41, 14.42
Gain margin 14.20, 14.21, 14.26
Gap 1.4,1.7, 1.19, 2.1, 2.6,2.10-2.13, 3.1,3.3, 3.6, 3.10, 3.20, 3.25, 3.74, 4.16, 5.44. 5.61,
5.62, 6.27, 6.28, 8.10, 9.5, 13.6, 13.16, 13.22, 13.25
Gauss 3.23, 4.7, 4.8, 13.19, 16.12
Cear 3.5, 14.7, 14.14
Geaimotor 11.2, 11.10
Generator 3.15, 3.71, 4.18, 6.39-6.41, 6.47, 7.4, 7.5, 7.7, 10.34, 14.17
Glassfibre 3.71, 4.28
"Go" coilside 3.42
Gramme 2.12, 5.65
Hague 5.64-5.66, 8.3, 8.20
Hall 1.19, 1.22, 1.24, 1.25, 2.45, 3.4, 3.16, 3.17, 3.30, 7.2, 11.7, 11.11, 11.12, 13.16
Hall effect 1.22, 3.16
Hall switch 1.22, 1.25, 3.4
Hard chopping 2.21
Hard disc 13.1
"Hard magnet 3.20
Hanmonic 1.11, 2.13, 2.23, 2.38-2.39, 3.2, 3.4, 3.45, 3.51,4.26, 4.27, 4.33, 5.8, 5.9, 5.26,
5.28-5.30, 5.33, 5.55, 5.61-5.65, 6.1-6.2, 6.9-6.10, 6.16-6.22, 7.13, 7.14,
8.2, 9.2, 11.13-11.14, 12.10, 12.24, 12;27, 13.8, 13.25
Heat 1.1, 1.17, 2,9, 2.13, 2.40, 3.28, S.32, 3.33, 3.73, 5.61, 6.42, 11.7, 11.8, 12.5, 12.6,
15.1, 15.2, 15.4-15.13, 15.22-15.24
Heat transfer 3.73, 12.5, 15.1. 15,5-15,8. 15.12, 15.22-15.24
Heat treatment 3.33
Henry 3.15
Hi-pot 3.31
High-energy magnet 2.3, 2.5, 5.61
Hoists 1.12
Hub 1.4, 4.2, 13.16, 13.18, 13.24
Hysteresis 1.17, 2.17, 2.22, 2.23, 3.25, 3.33, 4.2, 4.6, 4.7, 9.1. 10.13, 10.15, 10.21, 11.6,
12.27, 13.14, 14.14, 14.45, 16.1-16.3, 16.11-16.13, 16.16
Hysteresis brake 11.6
Hysteresis losses 3.25, 9.1, 12.27, 13.14, 16.11
IC 1.23, 2.30, 2.32, 2.46, 5.14, 5.24, 6.42, 6.44-6.46, 10.4, 10.16
Ideal drive 17.2
Identification 14.16
IEEE488 3.2, 11.15
IGBT 2.19, 2.20
Incremental encoder 1.22, 2.45
Incremental motion control 14.14, 14.15
Incremental permeability 4.2
Induced current 4.29
Inductance 1.17, 2.13, 2.16, 2.22, 2.32, 2.34, 2.43, 3.1, 3.6, 3.15-3.17, 3.33, 3.51, 3.71,
4.30, 5.13, 5.14, 5.25, 5.39-5.46, 5.48-5.51, 5.53-5.60, 5.62, 5.64. 5.65,

XXV
6.3, 6.4, 6.6-6.9, 6.13, 6.23, 10.9, 10.24, 10.25, 11.2-11.4, 11.6, 11.9,
12.19, 12.20, 12.23, 12.24, 13.2, 14.5, 14.7
Induction motor 1.1,1.2, 1.4, 1.12, 2-2, 2.6, 2.20, 2.46, 3.14, 5.29, 3.31, 3.35, 5.55, 5.58,
9.2, 13.1-13.3, 13.6, 13.7, 13.14, 13.15, 17.3
Inertia 1.4,1.19, 2.1, 2.2, 2.5, 3.2, 3.3, 3.6, 3.73, 6.46,11.2,13.1.13.2, 13.16, 14.5, 14.7,
14.8, 14.10, 14.14, 14.18
Infinite maximum speed 17.11
Inset 2.3
Inside rotor
see interior rotor
Instantaneous 1.9, 2.23, 2.34, 2.35, 2.46, 5.24, 5.32, 7.10, 7.12, 10.2
Insulation 1.16, 3.2, 3.28, 3.31, 3.32, 3.35, 3.70, 4.2,13.13,13,23,13.26,15.1,15.1,15.5,
15.10, 15.13, 16.15, 16,17
Integral gain 2.43, 14.22, 14.31
Integral slot 3.40
Integral-slot 3.10, 3.12, 3.34, 3.40, 3.48, 13.7
Interconnections 3.31
Interference 1.23, 2.39, 14.18, 14.32
Interior permanent-magnet motor
see IPM
Interior rotor 3.1, 3.3
Interior-rotor motor 1.4, 2.1, 2.2, 2.12, 4.11, 13.2
Interphase fault 3.36
Intrinsic 3.26, 3.27, 16.1, 16.2, 16.7, 16.8
Intrinsic coercivity 3.27, 16.2, 16.7, 16.8
Intrinsic demagnetization curve 3.26, 3.27
Inverter 1.10, 1.19, 2.39-2.41, 3.4, 4-33, 5.32, 7.2, 11.1, 11.6, 11.10, 15.4, 13.11, 13.15,
13.16, 16.13
Inverter utilization 17.2
IPM 17.8 et <*q., 18.1 et teq.
IPM parameter plane 17.16 et vq.
Iron 1.1, 1.17,1.20,1.22, 2.2, 2.3, 2.5,2.8, 2.11, 2.13,2.47, 3.4-S.6, 3.16, 3.24, 3.25, 3.28,
3.34, 4.2-4.5, 4.9, 4.18, 5.61, 5.66, 6.3, 8.3. 8,20, 9.3, 9.9, 13.15, 13.17,
13.21, 13,24, 15.13, 15.14, 16.6, 16.11, 16.13, 16.15, 16.1
Iron losses 2.13, 3.6, 8,3, 8.20, 9.9, 16.16
Kalman filter 14.46
Kapton 15.13
kE 1.13, 1.14, 3.18-3.20, 3.23, 5.12, 5.35, 5.36, 5.38, 5.39, 5.42, 7.1-7.10, 7.12-7.16, 8.1,
10.30, 10.33, 13.11, 13.13, 13.19, 13.25, 14.5-14.7, 14.9,
Keeper 4.4, 4.5, 16.1, 16.9
Kelvin 11.15, 15.6
Kevlar 1.4, 2.3, 13.6, 16.10
loj. 1.13, 1.14, 1.17, 3.19, 3.20. 3.22, 3.23, 5.12, 5.36, 5.38, 5.42, 6.39, 6.42, 7.1-7.10,
7.12, 7.13, 7.15, 7.16, 8.1, 11.8, 11.9, 12.21, 13.11, 13-25, 14.5
Lacing 3.35
Lamination 1.17, 2.2-2.S, 2.5-2.6, 2.8, 2.12, 3.1, 3.3, S.4-3.5, 3.9-3.10, 3.12, 3.24,
3.28-3.35, 3.50, 3.70, 4.2, 5.62, 8.11, 9.1, 9.6, 12;1. 12.6, 12.15, 12.16,
13.2-13.4, 13.7,13.9,13.15, 13.19,13.20,13.22-13.25,15,1,15.5, 15;10-
15.11, 16.11-16.12, 16.13, 16.15-16.16

xxvi
INDEX
Lamination stack
bond 3.26
Lap winding 3.10, 3.31, 3.35, 3.39, 3.40, 3.43, 5.9, 5.10, 6.3, 6.17, 6.19, 12.9, 1S.6, 13.8
Laplace operator 14.2
Laplace transform 14.1-14.3, 14.8, 14.18, 14.29, 14.36, 14.38
Laser S.28, 3.73, 13.23
Lead/lag 14.1, 14.25
Leakage 1.22. 2.3, 2.6, 2.11, 2.18, 3.8, 3.18, 5.21, 3.71. 4.3, 4.104.12, 4.14, 4.16, 4.2S,
4.25, 5.14. 5.25, 5.42, 5.43, 5.55, 5.57-5.59, 5.62, 6.13, 6.23, 6.29,
6.31-6.34. 6.41, 8.16, 12.18, 12,19, 13.9, 13.22, 13.25, 16.8
Leakage flux 1.22, 3.71, 4.10, 4.14, 4.16. 4.23, 5.43, 5.55, 8.16, 12.18, 12.19
Leakage inductance 5.14, 5.25, 5.55, 5.57, 5.62, 6.13, 6.23
Leakage permeance 4.3, 4.11, 4.12, 4.23, 4.25, 6.29, 6.33, 6.34
Leakage reactance 5.58, 6.23, 6.31, 6.32, 6.41
Line 1.3, 1.15, 1.18, 2.26, 2.27 , 2.29-2.35, 2.37-2.40, 2.45, 2.46,3.12,3.21.3.23,3.2
3.27, 3.29, 3.33, 3.45-3.48, 3.71, 4.7, 4.8, 4.22-4.24,4.26, 5.5,5.B, 5.14,
5.17-5.21, 5.24, 5.25, 5.29, 5.30, 5.32, 5.35, 5.37, 5.38, 5.43, 5.44, 6.4,
6.16, 6.29, 6.39, 6.41, 6.47, 7.2, 7.S-7.7, 7.9, 7.10, 7.13, 7.16. 8.1, 8.13,
10.2, 10.4. 10.7-10.11, 10.13, 10.15, 10.16, 10.18, 10.21, 10.22. 10.27,
10.28, 10.30, 10.31, 11.2, 11.4, 11.6, 11.9, 11.10, 11.14, 12.10, 12.25,
12.27,12.28,13.2,13.4,13.10,13.13,13.15. 13.19,13.22,13.26, 14.1
14.14, 14.31, 14.32, 15.9, 16.4
Line-line 2.31-2.33,2.38, 2.39,3.45-3.48, 3.71,5.14,5.17-5.21,5.24,5.30,5.35, 6.16, 6.41,
7.5, 7.7, 7.10, 7.13, 7.16, 8.1, 10.10, 10.27, 10.30, 11,4, 12.10, 13.13,
13.26
Linear 1.2, 1.11, 1.12, 1.15. 2.15, 2.38. 2.39, 4.16, 4.17, 4.19, 5.18, 5.36, 6.42, 8.5, 11.6,
12.2. 12.13, 12.14, 12.17, 12.19, 12.21, 14.1, 14.7, 14.16, 14.20, 14.34,
14.43, 15.12, 16.3, 16.4, 16.7, 16.14
Linearity 1.11, 2.41, 3.74, 5.38, 11.2, 11.8-11.10, 15.12
Liquid cooling 1.17, 2.13, 3.33, 15.8
Load 1.12, 1.14-1.16, 1.18, 1.19, 2.1, 2.6, 2.16, 2.17, 2.20-2.26, 2.39, 2.44, 3.15, 3.21,
3.23, 3.27, 3.33, 4.1, 4.7, 4.8, 4.18, 4.22, 4.33, 5.39, 5.66, 6.36, 6.41,
6.46. 7.4, 8.19, 10.1, 10.27, 10.28. 10.30-10.35. 11.2, 11.5-11.7, 11.9,
12.20, 12.21, 13.4, 13.11, 13.15, 13.19, 14.1, 14.5, 14.7-14.10, 14.14.
14.30, 14.31, 14.41, 15.7, 15.8, 15.17, 15.20, 16.8, 16.10,
Load line 3.21, 3.23, 3.27, 4.7, 4.8, 13.4, 13;15, 13.19
Locked rotor 1,14-1.16, 4.18-4,22. 11.7, 12.26, 13.15
Locate 2.8, 3.28, 3.29, 16,10
Lorentz 3.16, 3.17, 12.20
Losses 1.1, 1.4, 1.9, 1.17, 2.1, 2.6, 2.9, 2.11-2.13, 2.18, 2.19, S.5-3.6, 3.10, 3.25, 3.28,
3.32-3.34,4.2, 4;8. 4.284.33, 5.25, 5.26, 5.31, 5.39, 5.61, 6.41,6.46, 7.2,
8.3-8.4, 8.10, 8.20, 9.1-9.S. 9.5-9.9, 10.22, 10.30, 10.31. 10.34, 11.7,
11.12,12.20,12.21,12;27,13.7,1S.11,13.14,13.15,13.19, 13.22,13;2S,
13.24, 13.26, 14; 20, 15.1, 15.2, 15.4, 15.8, 15.9, 15.10, 15.12, 15.17,
16.11, 16.12-16.16
Low<arbon steel 2.2, 13.21
Machine tools 1.11, 1.12, 3.3, 3.76, 14.14

xxvii
Magnequench 13.19, 13.20
Magnet 1.1, 1.2, 1.4, 1.7, 1.9, 1.11, 1.16-1.17, 1.20, 1.22, 1.23, 1.25, 2.1-2.8, 2.10-2.15,
2.19-2.20, 2.40, 2.41, 2.46, 2.47, S.l-3.8, 3.15, 3.16-3.21, 3.23-3.27,
3.30-3.35,3.47, 3.50,3.51, 3.71,3.73-3.74,3.76,4.1-4.20,4.22,4.25-4.29,
4.32, 4.33, 5.1, 5.3-S.5, 5.8, 5.11-5.14, 5.17, 5.19-5.21, 5.24-5.26,
5.28-5.32, 5.35, 5.38-5.39, 5.44, 5.52, 5.55, 5.61-5.66, 6.1-6.4, 6.9, 6.10,
6.12, 6.16. 6.21, 6.22, 6.24, 6.26, 6.28. 6.29. 6.31, 6.33-6.35, 6.37-6.40,
6.42-6.44, 6.46-6.48, 7.10-7.16, 8.2-8.13, 8.15, 8.16, 8.19, 8.20, 9.5, 9.9,
10.15,11.4,11.5,11.7-11.9,11.11,11.12,11.16,12.6,12.9-12.10, 12;15,
12.17, 12.18, 12.20-12.28, 13.1, 13.2, 13.4-13.11, 13.15-13.22, 13.25-
13.26, 14.1, 14.41, 15.13-15.14, 16.1-16.12, 16.16, 16.17
bonded ring 3.8
embedded slab 3.8
ferrite 3.24
Magnet length 3.21, 4.12, 4.14, 8.2, 8.6, 13.22, 16.8
Magnet thickness 3.24-3.26,4.18,12.6,12.22,13.4,13.6,13.9,13.10,13.16,13.19.13.20
Magnet wire 3.30-3.32, 13.18, 16.17
Magnetic 1.1,1.2,1.4,1.6,1.7,1.14,1.17,1.19,1.22, 2.1, 2.3, 2.6, 2.11-2.13, 3.1, 3.6, 3.8,
3.15, 3.16-3.18, 3.21, 3.26, 3.29, 3.32, 3.33, 4.1-4.6, 4.8, 4.9, 4.11-4.13.
4.15-4.18, 4.23-4.29, 4.32, 5.1, 5.32, 5.36, 5.38, 5.39, 5.42-5.44, 5.52,
5.61, 5.63-5.65, 6.9, 6.12. 6.24, 6.25, 6.28, 6.34, 6.36, 8.2-8.4, 8.9, 8.16,
8.20,9.1, 9.2,10.2,10.34,11.6,11.9,12.2-12.4,12.6, 12.13,12.19-12.21.
12.28, 13.3, 13.4, 13.6, 13.17, 13.19, 15.22, 14.7, 14.9, 16.1, 16.8-16.13,
16.15, 16.16
Magnetic loading 2.12, 3.1, 12.2-12.4, 12.6
Magnetic short circuit 16.1
Magnetization 3.20, 3.21, 3.25, 4.6, 4.9, 4.12, 4.28, 6.2, 6.3, 11.11, 11.12, 12.18, 12.21,
12.25, 13.7, 13.9, 16.1, 16.3, 16.7. 16.8. 16.11, 16.12, 16.
Magnetizing fixture 3.8, 4.5, 13.9, 16.8
Maximum energy product 4.15, 16.4, 16.7
Maxwells equations 12.13
Mean length of turn 5.60, 13.13
MGOe 5.23, 4.8, 4.15
Micro-controller 14.11
MLT 3.70, 5.60, 13.13, 13.26
Model reference adaptive control 14.43, 14.45
Modulation 1.17, 2.15, 2.21, 2.22, 2.38, 2.39, 2.44, 6.2, 9.1
Moisture 3.32
MOSFET 2.19, 2.20
Motion control 3.76, 14.14, 14.15, 15.14
Motor 1.1-1.25, 2.1-2.3, 2.5, 2.6-2.20, 2.22, 2.23, 2.27, 2.28, 2.30, 2.32, 2.34, 2.37-2.47,
5.1-3.6, 3.8-3.19, 3.21, 3.23, 3.24, 3.26-3.36, 3.39-3.48, 3.50, 3.51,
3.71-3.76, 4.1-4.5, 4.9-4.11, 4.1M.16, 4.1&4.20, 4.22, 4.23, 4.26, 4.28-
4.29, 4.32, 4.33, 5.1-5.4, 5.6-5.12, 5.14-5.16, 5.18-5.26, 5.28-5.33,
5.35-5.41,5.43, 5.44,5.52, 5.55, 5.57, 5.58, 5.60-5.63, 5.66, 6.1-6.4, 6.11,
6.12, 6.16, 6.23, 6.25, 6.29, 6.35, 6.37, 6.39, 6.40, 6.42-6.44, 6.46, 6.46,
6.47, 7.1-7.16, 8.1-8.3, 8.5, 8.6, 8.10. 8.12, 8.19, 8.20, 9.1, 9.2, 9.9,
10.1-10.6, 10.9, 10.16, 10.18, 10.20, 10.22, 10.27, 10.28, 10.30-10.33,
10.35, 11.1, 11.2,11.4-11.14, 11.16,12.1, 12.2-12.9, 12.13-12.21, 12.24,

xxviii
I n d ex
13.1-13.4,13.6-13.11,13.14-13.16,13.18,13.19,13.22,13.24-13.26,14.1,
14.5, 14.7-14.10, 14.12-14.19, 14.28, 14.31, 14.32, 14.34, 14.40-14.46,
15.1, 15.2, 15.4, 15.5, 15.7-15.17, 15.19, 15.21-15.22, 16.1, 16.4,
16.6-16.16
AC induction 1.1
AC servo 1.2
stepping 1.1
velocity servo 11.9
Motor-in-hub 18.16, 13.18, 13.24
Motoring 2.44, 3.2, 3.15, 6.37, 10.1, 10.34, 10.35, 14.12
Muitiple-strand 3.32, 16.16
Mutual inductance 3.15, 3.71, 5.13, 5.42-5.44, 5.46, 5.48-5.51, 5.53-5.55, 5.57-5.59, 5.65,
6.8, 6.9, 11.4, 12.23
NdFeB
see Neodymium Iron Boron
Needle 3.12, 3.31, 5.21, 11.12, 13.18
Negative feedback 14.11, 14.31
Neodymium-Iron-Boron 3.24, 3.27, 4.3, 4.5, 4.9, 11.7, 12.4, 13.19, 15.13, 16.6
Nested 3.29, 3.70, 14.32
Noise 1.1,1.19,1.22,1.25,3.2, 3.34,12.1,13.18,13.22,14.18,14.31,14.32,14.34,14.41,
14.46
audible 1.19
Nomex 15.13
Nonlinear 2.37, 4.16, 4.25, 8.16, 11.8, 12.19, 14.7, 14.20, 14.41, 14.43, 15.1
Non-salient 17.6
Nyquisi 14.35
Observer 6.12, 6.23, 14.46
OD 2.3, 3.35, 6.44, 12.5, 12.6, 13.3, 13.6, 13.9, 13.20, 13.21
Oflset 3.5, 3.48, 6.42
Op-amp 2.43, 14.11
Open-circuit 3.15, 3.20, 3.21, 3.23, 4.5, 4.7, 4.8, 4.12, 4.14-4.16, 4.19, 6.3, 6.12, 6.22,
6.33, 6.35, 6.45, 7.4, 11.5, 12.21, 12.22, 12.26, 16.8
Open-loop transfer function 14.21, 14.22
Operating region 1.16, 1.18
Optical interrupter 1.25
Optimal control 14.1, 14.46
Oscilloscope 3.19, 7.7, 11.2, 11.5, 11.11, 11.14
Output 1.17,1.19, 2.16, 2.39, 3.3, 3.5, 3.51,4.18,6.41,11.1,11.2,11.7,11.11,12.2,12.5,
12.7, 12.10, 12.13, 12.19, 13.14, 13.19, 14.2, 14.5, 14.10-14.13, 14.16,
14.20-14.22,14.26,14.31,14.33,14.34,14.36,14.40,14.41,14.43,14.45,
14.46, 15.8, 15.15
Output coefficient 12.2
Output equation 12.2, 14.36
Outside rotor 3.25
see exterior rotor
Overcurrent 2.16, 2.30, 2.41, 3.2, 4.28, 5.38, 6.46, 6.47
Overload factor 15.15-15.20

x xix
Overtemperature S.2, 4.28, 5.38
Pack 3.28, 3.30, 3.31
Pancake motor 2.1, 2.10, 2.11
Pancake resolver 1.20
Parallel 2.19, 3.18, 3.26, 3.71. 4.3, 4.12, 4.204.22, 5.12. 5.24, 5.35, 5.39, 5.46, 5.47. 5.59,
5.60, 5.63, 5.64, 6.2, 6.7, 6.14, 6.28, 6.29, 6.33, 6.35, 7.12, 7.13, 7.15,
7.16,10.13, 10.15, 10.16,10.20,10.21,13.11, 13.15,14.12, 15.7,1
15.15, 16.10
Parallel paths 3.18, 3.26, 4.20, 4.21, 5.12, 5.35, 5.39, 5.46. 5.47, 5.59, 6.7, 6.14, 6.28,
7.12, 7.13, 7.15, 7.16, 13.11, 13.15
PC-BDC 5.13, 5.30, 5.51,5.59, 8.1, 8.3, 8.13.8.17, 9.2, 9.7,10.1,10.26,12.6-12.14,13.27,
14.41, 15.23
PC-SRD 9.2
Performance 1.2, 1.5, 1.12. 2.10, 2.11. 2.13, 2.20, 2.43, 3.1, 3.3, 3.17, 3.22, 3.28, 4.9,
5.33, 5.66, 6.1, 6.3, 6.42, 8.3, 8.20, 11.1, 11.2, 11.7, 12.1, 12.4,
12.7-12.10, 12.14, 13.1, 13.19, 13.23, 13.26, 14.1. 14.9, 14.10, 14.12,
14.18, 14.25, 14.26, 14.40, 14.43, 16.4, 16.6, 16.11, 16.15
Permanent magnets 1.1
Permeability 4.2, 4.6, 4.7, 4.18, 5.36, 8.6, 8.16, 11.4, 12.20, 12.22, 16.2, 16.4, 16.7, 16.11
Pcrmeance 2.3,3.20, 3.23. 3.24,4.3, 4.8,4.11, 4.12,4.14, 4.15,4.23-4.27, 5.13, 5.55-5.59,
6.24, 6.28, 6.29, 6.33, 6.34, 13.4, 16.4, 16.7, 16.8
Phase 1.2, 1.9-1.11, 1.13, 1.17-1.19, 1.21-1.23, 2.1, 2.2, 2.12, 2.14-2.17, 2.19, 2.23-2.31,
2.35-2.41,2.44-2.47, 3.1,3.3-S.5,3.9-3.13,3.18-3.20,3.26-3.28, 3.31-3.36.
3.38,3.41-3.51,3.70.3.71,4.1,4.5,4.7-4.10,4.12,4.15,4.184.22,5.1-5.3,
5.5, 5.6, 5.9, 5.11-5.14, 5.17-5.21, 5.23-5.26, 5.2&-5.30, 5.32-5.35,
5.38-5.41, 5.43, 5.44, 5.46-5.50, 5.54, 5.55, 5.58-5.60, 6.2-6.4, 6.7, 6.8,
6.11-6.14, 6.16, 6.19-6,23, 6.27, 6.28, 6.31, 6.32, 6.35, 6.38, 6.41. 6.43.
6.46, 6.47, 7.1, 7.4, 7.6-7.16, 8.1, 8.9, 8.10, 8.16, 8.17, 10.1, 10.3, 10.4,
10.8,10.12,10.15,10.16,10.22-10.26,10.28,11.2-11.7,11.9,11.14,12.3.
12.10, 12.13, 12.22, 12.24, 13.2, 13.6, 13.8, 13.11, 13.13-13.15, 13.25,
13.26, 14.9,14.16,14.18,14.20, 14.21,14.25,14.26,15.14
Phase displacement 3.42, 3.43, 3.45, 5.5
Phase margin 14.20, 14.21, 14.26
Phase sequence 10.4
Phaseleg 2.15-2.17, 2.26, 5.32
Phasor 2.38, 2.46, 6.3. 6.4. 6.22, 6.28, 6,30, 6.32, 6.35-6.37, 6.3M.46, 7.15, 11.4, 11.5,
12.24
Phasor diagram 2.38, 2.46, 6.3, 6.4, 6.22, 6.28, 6.30, 6.32, 6.35-6.37, 6.39-6.42, 6.44-6.46,
11.4, 11.5, 12.24
PID control 14.30, 14.31, 14.41, 14.43
Pitch 1.21, 3.10, 3.12, 3.14, 3.31, 3.35, 3.36, 3.40, 3.42, 3.45, 3.48. 3.50. 3.51, 4.12, 4.17,
4.30, 4.32, 5.2, 5.3, 5.5, 5.6, 5.8, 5.9, 5.11, 5.14, 5.20. 5.28-5.30, 5.35,
5.36, 5.44, 5.45, 5.53, 5.58, 6.1, 6.3. 6.6, 6.7, 6.1^6.17. 6.19-6.21. 7.11,
7.12, 7.14. 8.1, 8.4, 8.5, 8.9-8.12, 8.17. 10.15, 12.5, 12.16-12.18,
13.6-13.9, 13.13, 13.18, 13.20, 13.22, 13.25
Pitch factor 3.45, 5.29, 6.16, 6.17, 6.21
Plant 14.7, 14.11, 14.12, 14.14, 14.20, 14.26, 14.30. 14.31, 14.34, 14.36. 14.39, 14.40,
14.43
Platen 13.16

XXX
Index
Polarization 16.1, 16.2
Pole-pieces 2.5, 3;7, 3.24, 3.74, 4.23, 4.25, 6.3. 6.2W.29, 6.33
Poles 1.1, 1.2, 1.23, 2.3, 2.5, 2.6, 2.10, 3.1, 5.6-3.13, 3.20, 3.23, 3.25, 3.33-3.36, 3.41,
3.43-3.45, 3.48, 3.50,3.71,4.4,4.5,4.21.4.27.4.28,4.32, 5.5,5.12,5.17,
5.18, 5.21, 5.35, 5.46. 6.4, 6.7, 6.1S, 6.25, 8.6. 8.7, 8.16, 12.21, 13.1,
13.6-13.10, 13.18, 13.21, 13.22, 13.26, 14.5, 14.8, 14.12, 14.16,
14.22-14.26, 14.30, 14.38, 16.15
Position control 1.1, 1.11, 2.45, 14.13-14.15, 14.18, 14.28-14.30. 14.32
Power 1.2.1.9, 1.10,1.12, 1.15-1.17, 1.19, 1.25, 2.3, 2.5, 2.6, 2.11, 2.14-2.24, 2.27, 2.80,
2.39-2.42, 2.44, 2.45, 2.47, S.2-3.6, 3.15, 3.25, 3.26, 3.28, 3.32, 5.1, 5.5,
5.17.5.32, 5.33, 5.39, 6.4,6.38, 6.44, 6.46,6.47, 7.4,7.7-7.10, 7.15, 8.19,
9.1, 9.9, 10.1, 10.2, 10.30, 10.31, 10.34, 11.1, 11.2, 11.6. 11.7, 11.13.
11.14, 12.3, 13.13, 13.14, 13.18, 13.19, 14.2. 14.18. 14.46, 15.1, 15.5,
15.8, 15.15, 15.17, 15.24, 16.8, 16.16
Power density 2.6, 3.28, 13.18, 15.1, 15.5
Power factor 2.6, 2.20, 2.39, 6.38, 6.44, 6.46, 7.15
Power transistor 1.2, 1-10, 1.16, 1.19, 2.14, 2.15, 2.18, 2.19, 2.21, 2.23, 2.30, 2.41, 2.42,
3.4, 5.5, 5.17, 5.32, 5.33. 5.39. 10.1
PPR 14.13
Precision dynamometer 11.12-11.16
Prime mover 10.34, 16.10
Principle of virtual work 3.17
Progressive die 3.28
Properties 2.18, 4.6, 4.9, 4.12, 6.1, 6.4, 7.10, 13.4, 14.7, 14.12, 15.1, 15.6, 15.13, 16.1,
16.3-16.5, 16.7, 16.8, 16.11, 16.12, 16.15, 16.17
Proportional gain 2.42, 14.31, 14.41, 14.42
Pulse-width modulation 1.17, 2.15, 6.2, 9.1
Pulses per revolution (PPR) 14.13
Punch press 3.28
Punching 2.3, 3.28, 3.29, 4.27, 9.1, 9.2, 13.2, 13.23-13.25, 16.15, 16.16
PWM 2.21, 2.23, 2.30, 2.35, 2.38, 2.39, 2.41, 2.43-2.46, 3.25, 4.33, 5.32, 6.2, 6.43, 7.2,
10.10, 10.28, 10.33, 11.12, 14.34. 16.11, 16.13
q-axis 2.46, 4.23. 6.2, 6.3, 6.9, 6.22-6.24, 6.28, 6.29, 6.31-6.33. 6.35-6.38, 6.40, 6.42-6.44,
6.46, 7.15. 8.5, 11.4, 12.21. 12.26
Quadrant 2.17, 4.6-4.9, 4.16, 5.64, 6.38, 14.1, 14.12, 14.13, 16.3, 16.4
Quadrants 2.44, 14.12, 16.16
Radiation 15.1, 15.2, 15.5, 15.8-15.12
Rated currcnt 17.1 et xeq.
Rated speed 17.1 rt vq.
Rated torque 17.1 et vq.
Rated voltage 17.1 et vq.
Recoil line 4.8, 16.4
Recoil permeability 4.6, 4.7, 8.6, 8.16, 12.22, 16.2, 16.4, 16.7
Rectifier 2.17, 2.39, 2.40, 2.44, 6.41, 6.47, 7.5, 7.7, 10.28
Reference 1.7, 1.21, 1.25, 2.15, 2.22, 2.23, 2.38, 2.41, 2.43. 2.46, 2.47, 3.1, 3.19, 3.45,
4.32, 5.1, 5.2, 6.2, 6.10, 6.35, 7.2, 7.9, 7.10, 10.35, 12.28, 13.6, 13.8
13.25, 14.11-14.13, 14.22, 14.43, 14.45, 15.24, 16.16. 16.17

XXXI
Reference frame 2.23, 2.46, 7.10
Reference signal 2.43, 14.11-14.13
Regenerative 2.44, 2.45
Regenerative braking 2.44, 2.45
Regulation 2.20-2.2S, 2.S4. 2.39, 2.42, 2.43, 2.45. 5.32, 10.10, 10.18, 10.24, 10.28
Regulator 2.15, 2.23, 2.38, 2.46, 10.9, 10.13, 10.14, 10.18. 10.21, 10.23, 10.24. 10.27,
14.43,14.45
Reluctance 1.12, 1.19.1.25, 2.3, 2.13, 4.3, 4.12, 4.24, 5.55, 5.66, 6.29, 6.36, 6.38, 7.16,
8.2, 8.4, 8.16, 10.35, 12.22
Reluctance network 8.16
Remanence 3.23-3.25, 13.4, 16.1-16.3, 16.7, 16.8
Remanent 3.21, 4.S4.9, 4.13-4.15, 11.9, 11.12. 16.11
Remote sensing 1.23, 2.47, 13.18
Resistivity 1.4, 1.17, 3.33, 4.32, 15.2, 15.4, 15.12, 16.11. 16.17
Resistor 2.44, 2.45, 6.47, 10.24
Resolution 1.19, 1.21, 1.25. 11.13, 14.13, 14.14, 14.35
Resolver 1.19-1.22, 1.25, 2.45, 6.11, 14.10, 14.13, 14.34
Resolver-todigital converter 1.19, 1.25, 14.34
Resonance 1.11, 14.5, 14.9
Retaining can 1.4, 2.3, 4.28-4.30, 6.41
Retention 2.1-2.3, 2.8. 3.24, 13.4, 13.6
Retrogressive 3.42
"Return" coilside 3.42
Reversal 1.2, 5.17
Reverse 1.2, 2.17-2.19, 2.24, 2.25, 2.43, 2.44, 2.46, 3.2, 5.4, 4.8, 4.22, 5.3, 6.29, 10.9,
10.15, 10.27, 10.28, 14.1
Reverse commutation 2.43
Reverse recovery 10.9
Ripple 1.11, 1.17-1.19, 2.6, 2.13, 2.22, 2.23, 2.30, 2.46, 2.47, 2.49, 3.5, 3.6, 3.19, 3.50,
3.51, 4.26-4.28, 4.33, 5.19, 5.20, 5.25, 5.26, 5.29-5.31, 5,62, 5.63, 6.21,
11.2, 11.10, 11.11, 14.1
Robust 12.14, 14.10, 14.26, 14.30, 14.40, 14.43, 14.46
Robustness 14.1, 14.11, 14.12, 14.25, 14.30, 14.43, 14.46
Root locus 14.22-14.26, 14.38
Rotation 1.19, 2.3, 2.11, 2.43, 2.45, 3.4-3 6, 3.15, 3.31, 4.26, 4.27, 4.31. 5.32, 5.41, 6.2,
6.10, 6.11, 8.9, 8.11, 9.2, 11.7, 12.16. 13.7, 13.16, 13.22
Rotational voltage 3.15, 3.16
Rotor 1.1-1.7,1,9,1.14-1.16,1.19,1.20,1.22-1.24,2.1-2.6. 2.8-2.10, 2.12, 2.20, 2.30, 2.32,
2.41, 2.4S, 2.44, 2.46, 3.1, 3.3, 3.4, S.6-3.8, 3.10, 3.17, 3.18, 3.20,
3.24-3.26.3.28,3.33-3.36,3.51,3.71,3.73-3.75,4.2,4.11,4,12,4.17-4.23,
4.25-4.31,4.33,5.1-5.3,5.14.5.19,5.21,5.25,5.31,5.33, 5.
5.55,5.61,6.2,6.9-6.12,6.23, 6.25,6.28.6.29,6.33-6.35,6.41, 6.44,7.11,
7.12, 8.3, 8.4, 8.9, 8.11, 8.13, 8.15, 8.16, 9.2, 9.3, 9.6, 9.7, 10.2. 10.24,
11.1, 11.2, 11.4, 11.7, 11.9, 11.12, 12.2-12.7, 12.12, 12.15-12.17,
12.19-12.21, 12.23-12.26, 13,2, 13.4, 13.6-13.11, 13.15, 13.16,
13.18-13.23, 13.26, 14.5, 15.2, 15.9, 15.11, 16.9, 16.10. 16.13, 16,15
axial 3.6
inside 1.4, 3.6
outside 1.4, 3.6
In d ex
Roving 3.71, 4.28
Saliency ratio 18.4
S.I. Units 7.1
Samarium cobait 3.8
Sampling rate 14.35. 14.36, 14.41
Screening 14.18
Second-order system 14.24, 14.28, 14.29
Self-inductance 3.15, 3.71, 5.42-5.46, 5.48-5.51, 5.55. 5.59, 5.65, 6.7, 6.9, 11.4, 12.23
Self tuning regulators 14.43
Sensor 1.20, 1.22, 2.30, 2.33-2.35, 3.5, 3.71, 5.14, 6.11, 7.7, 11.2, 11.6, 11.7, 14.13
Sensorless control 1.23
Separately-excited DC motor 17.1
Sequence 1.22, 1.23, 1.25, 2.27, 2.32, 2.33, 2.35, 3.43, 3.44, 5.5, 5.14, 5.15, 5.21, 5.23,
5.26, 5.28-5.31, 10.2, 10.4, 10.13, 10.20, 10.24, 11.9, 14.36, 15.14, 16.4
Servo 1.2, 1.4, 1.11, 1.12, 1.24. 2.1, 2.5, 2.41, 2.44, 2.49, 3.3. 3.4, 3.32. 3.76, 4.33, 5.38,
7.1, 13.1, 13.2, 13.26, 14.7, 14.9, 14.12, 14.18, 14.46,
Servo motor 1.12, 3.71, 3.72, 3.76, 4.26. 5.38, 11.9, 13.1, 13.2, 15.14, 15.17
Servo system 14.7, 14.9, 14.12
Settling rime 14.19, 14.25, 14.28, 14.29, 14.39
Shaft coupling 14.9
Shaft flux 3.8
Shaft position transducer 1.19, 1.22, 1.23, 1.25
Short-pitching 3.36, 3.51, 6.3, 6.16
Short-time operation 1.16, 15.22
Shunt 1.22
SI Units 3.19, 4.6, 15.10
Signal 1.19-1.22, 1.25, 2.11. 2.19, 2.22, 2.30, 2.41-2.45, 3.32, 7.2, 10.28, 14.2, 14.5.
14.10-14.18, 14.20-14.22, 14.31, 14.32, 14.34-14.36. 14.40, 14.41
Silicon 2.19, 2.47, 3.4, 3.5, 3.33, 4.2, 5.33, 13.23, 15.13. 16.11
Silicon steel 3.33. 13.23. 15.13. 16.11
Simulation 5.26, 5.29.5.30,5.58, 8.1, 10.1,10.11,10.28,10.31, 10.33,12.1,12.10,12.24,
14.9, 14.34, 14.41, 15.22
Sine 1.19,2.38, 2.39,3.4,5.47, 6.1-6.10,6.13, 6.14, 6.19,6.21-6.25. 6.27, 7.2, 7.14,12.24,
14.34
Sinewave 1.2, 1.3, 2.5, 2.15, 2.28-2.30, 2.37, 2.38, 2.40, 2.45-2.47. 3.4, 3.5, 3.36, 3.39,
3.45, 3.48, 3.50, 3.74, 4.26,4.30, 4.31. 5.1, 5.3, 5.9,5.14,5.18,5
5.24, 5.31, 5.33, 5.36, 5.39,5.58, 5.62, 6.1-6.4, 6.35,6.37,6.40,6
6.44, 7.1, 7.2, 7.7-7.10, 7.13-7.16,8.1, 9.1, 9.3, 11.1, 11.4, 11.5, 12.9.
12.10, 12.24, 13.8, 13.9
Sinewave drive 2.5, 2.37, 2.40, 2.45, 2.46, 3.5, 3.50, 3.74, 11.1, 11.4, 11.5, 12.9, 13.8
Single phase 2.26, 3.4, 5.32, 13.8, 13.11
Single-quadrant 14.1, 14.12
Skew 3.34, 3.35, 3.50, 3.51, 4.27, 4.28, 5.9, 5.29, 5.63, 6.3, 6.21, 7.14, 8.2, 8.3, 8.7, 8.8,
8.11, 8.14, 13.9, 13.11, 13.25
Sliding mode 14.43, 14.45
Slip rings 1.1
Slot 2.3, 2.6, 2.13, 3.1, 3.4, 3.6, 3.9-3.14, B;16, 3.20, 3.21, 3.29-3.32, 3.34-3.36, 3.38-3.45,

xxxiii
5.4651, 5.70, 5.71, 4.2, 4.264.30, 4.32, 5.2, 5.3, 5.5, 5.6, 5.9, 5.11, 5.12,
5.14, 5.20, 5.21, 5.31, 5.36, 5.39, 5.42-5.44, 5.47-5.51, 5.55-5.59, 5.61-
5.62, 6.13, 6.16-6.19, 6.21, 6.31, 6.32, 8.1-8.3, 8.9-8.12, 8.15, 8.17, 8.19,
10.31,11.12,12.4-12.6,12.9,12.16,12.18,12.22,12.23,15.2,15.3,15.6
15.8, 15.10-15.13, 15.18, 15.22, 15.23, 15.25, 15.26, 15.1, 15.5, 15.23,
15.24
Slot fill 5.70, 12.4, 12.5
Slot liner 5.30, 5.70, 15.1, 15.24
Slot opening 5.34, 4.30, 4.32, 8.1, 8.2, 8.12, 15.3, 15.13, 13.22
Slot-fill 3.70, 5.39
Slotless 1.4, 2.1, 2.12-2.14, 3.16, 5.1, 5.61-5.63, 8.3
Slots/pole 3.10, 5.12, 5.14, 3.34, 5.40, 5.41, 5.44, 3.45, 3.50, 4.264.28, 5.2, 5.11, 5.36,
5.48, 5.50, 8.19, 12.18, 15.7, 15.8
Slotting 1.17, 1.19, 2.1, 5.28, 4.2, 4.13, 5.21, 5.28, 5.29, 5.44, 5.62, 6.21, 8.3, 8.5, 8.9,
12.22, 16.8
Soft chopping 2.21, 2.23
Space harmonic 5.61, 6.16
Space-harmonic 2.13, 3.4. 4.27, 5.55. 5.61, 5.63, 5.65, 6.9, 6.18, 7.14, 12.24
Span 5.12, 5.14, 3.33, 3.35. 3.36. 3.39, 3.41-5.43, 3.45, 3.49, 3.51, 4.19, 5.3, 5.8, 5.20,
5.53, 6.3. 6.6, 6.16, 6.17, 6.20, 8.2
Specific heat 15.6, 15.10, 15.24
Speed 1.1, 1.2, 1.4, 1.9, 1.11-1.19, 1.21, 1.23, 1.25, 2.1, 2.3, 2.6, 2,10, 2.11, 2,13-2.15,
2.19, 2.34. 2.41-2.47, 5.2. 5.3, 3.6, 3.14, 3.18, 3.28, 5.29, 3.32-3.34,3.71,
4.1, 4.4, 4.18, 4.26, 4.28, 4.31-4.33, 5.39-5.42, 6.4, 6.10, 6.12, 6.23,
6.42-6.44, 6.46, 6.47, 7.S-7.5, 8.19, 9.9, 10.1, 10.2, 10.4, 10.9, 10.14,
10.18,10.25,10.27,10.28,10.30-10.35,11.1,11.2,11.5-11.7,11.9,11.10,
11.12,12.6,12.9,15.2,13.4,15.11,13.14,13.16,13.19,13.26,14.1,14.2,
14.5-14.14, 14.17-14.19, 14.31, 14.32, 14.34, 14.39, 14.46, 15.9. 15.23
Speed reference 2.43, 14.11-14.13
Speed/torque 1.2, 1.12-1.18, 3.71, 4.1, 4.4, 4.33, 5.39, 5.42, 6.4, 6.42, 6.44, 6.46, 7.3,
11.6, 11.7, 13.14, 14.12
see also Chapters 17 and 18
Spindle motor 2.1, 17.3
Spoke-iype motor 5.74, 5.75, 6.25-6.26, 6.28, 12.18, 12.19
Squarewave 1.2, 1.3,1.5. 1,10-1.12, 2.26-2.31, 2.35-2.37, 2.39, 2.40, 2.42, 2.43, 2.45-2.47,
3.5, 5.6, 5.18, 5.19, 5.39, 5.46, 5.47, 5.49-5.51, 5.71, 4.19, 5.1, 5.3, 5.5,
5.8, 5.14-5.16, 5.18, 5.19, 5.21-5.23, 5.25, 5.28, 5.33, 5.355.37, 5.41,
5.43,5.47, 5.62, 6.1, 6.4,6.12,6.35, 7.1, 7.2, 7.4,7.7,7.8,7.10-7.13, 7.16,
8.1, 10.1, 10.28, 11.4, 12.9, 12.10, 12.21, 12.24, 13.8, 15.13, 14.5
Squarewave drive 2.26-2.28, 2.35, 2.36, 2.40, 2.42, 2.43, 2.45, 2.46, 3.5, 5.6, 3.46, 5.47,
3.49, 3.50, 3.71, 7.16. 11.4, 13.8
Stability 2.42, 14.1, 14.12. 14.13, 14.15, 14.16, 14.20, 14.26, 14.31, 14.32, 14.38, 14.43,
14.46, 16.7
Stack 2.2, 2.5, 2.8, 2.12, 3.12, 3.24, 3.28-3.30, 3.32-5.35, 5.50, 5.70, 4.12,4.28, 4.32, 5.35,
7.11, 8.4, 8.7,12.2,13.2-15.4,13.9-13.11,13.13,15.22-13.25,15.4,15.5,
15.11
Stall torque 1.14, 3.76
Stamping 3.29, 4.2, 5.57, 13.25, 16.13, 16.15, 16.16
Star-delta switching 17.5

xxxiv
Index
State observer 14.46
State-space 10.10, 10.11. 10.18, 10.25, 14.45
Stator 1.1, 1.4,1.11,1.17,1.19, 1.20, 1.22, 2.1-2.3, 2.6, 2.8, 2.10-2.13, 2.46, 3.1, S.3, 3.4,
3.6, 3.8-3.10,3.14,3.20,3.21, 3.24. 3.26-3.29,3.31-3.36, 3.39, 3.40, 3.43,
3.70, 3.73, 4.1, 4.2, 4.12, 4.164.19, 4.26, 4.28, 4.30, 4.33, 5.2, 5.3, 5.5,
5.12, 5.14, 5.18, 5.20, 5.21, 5.24, 5.26, 5.31, 5.35, 5.36, 5.38, 5.44-5.46,
5.51,5.53-5.55,5.60-5.63,5.65,6.1,6.3, 6.11, 6.12, 6.29,6.31,6.35,6.37,
6.39, 7.11, 7.13, 7.14,8.M.6, 8.9-8.11, 9.1, 9.2, 9.5, 9.6,11.1, 11.2,11.4,
11.7, 11.9, 12.3-12.6, 12.15-12.17, 12.21, 12.26, 13.1-13.4, 13.&-13.8,
13.15, 13.16, 13.18, 13.19, 13.22, 13.23, 13.25, 15.4, 15.5, 15.9, 15.11
15.23, 16.10, 16.13, 16.16
Steady-state error 14.1, 14.21, 14.22, 14.25, 14.26, 14.31, 14.32, 14.39, 14.41, 14.43
Steel 1.4, 1.22, 2.2, 2.3, 2.6, 2.8, 2.10, 2.11, 2.13, 3.17, 3.21, 3.25, 3.33, 3,73, 4.1, 4.2,
4.12, 4.16-4.18, 4.31, 4.32, 5.44, 5.55, 5.60-5.62, 9.1, 11.12, 12.3, 12.4,
12.15,12.17, 13.6, 13.14, 13.15, 13.21, 13.23, 15.5, 15.13, 15.14. 15.23,
16.1, 16.9-16.11, 16.13-16.16
Stepper 1.1, 1.2, 1,19
Stepping motor 1.1
Summing junction 14.11, 14,34, 14.41
Surfece-magnet rotor 2.5, 3.7, 3;21, 3.23-3.25,4.2, 4.11,4.13,4.14,4.18, 4.28,5.44, 5.52,
5.55, 6.4, 6.26, 6.35, 6.39, 6.42, 6.43, 6.46, 8.5, 8.10, 8.12, 11.4
Switchmode 2.15, 2.23, 2.41
Synchronous 1.1, 1.2, 2.2, 2.3, 2.6, 2.15, 2.20, 2.46, 3.6, 5.14, 5.18, 5.66, 6.2, 6.3, 6.9.
6.10, 6.12, 6.23, 6.24, 6.32, 6.35, 6.36, 6.40-6.43, 6.46, 6.47, 7.8, 7.9,
11.2, 11.4, 11.5, 11.9, 12.20, 12.24, 12.28
Synchronous reactance 3.6, 6.3, 6.12, 6.23, 6.24, 6.32, 6.36, 6.41, 6.42, 6.46, 6.47,11.4,
11.5, 12.24
Synchronous reluctance 2.3, 17.26
System identification 14.16
Teeth 1.1, 2.6, 2.12, 2.13, 3.16, 3.24, 3.29, 3.33, 3.34, 4.2, 4.17, 4.27, 4.28, 5.36, 5.38,
5.55, 5.61, 6.29, 8.9, 8.10, 8.14, 8.16, 8.18, 9.2, 9.3, 9.6, 9.8, 12.3, 12.4,
12.21, 12.26, 12.27, 13.2. 13.12, 13.18, 13.22, 13.23. 16.15
TEFC 15.13
TENV 15.12, 15.13
Test facilities 11.12
Thermal capacity 11.8, 15.15
Thermal conductivity 15.3-15.6, 15.10, 15.11, 15,23, 15.24
Thermal resistance 4.2, 11.2, 11.7, 11.8, 13.14, 15.4, 15.5, 15.12, 15.15, 15.23
Thermal time constant 11.1, 15.22
Thermistor 11.7
Thermocouple 15.24
Three-phase 1.10,1.22-1;23.2.2,2.16, 2.27-2.31,2.35-2.39, 2.44, 2.45, 2.47, 3.4, 3.5, 3.26,
3.27,3.35,3.41, 3.44, 3.46-3.48,3.50, 3.71, 5.1, 5.3,5.5, 5.14, 5.18, 5.20-
5.21, 5.32, 5.33, 5.43, 5.44, 5.58, 6.11, 6.13, 6.16, 6.47, 7.1, 7.4, 7.8, 7,9,
7.11, 7.15, 10.28, 11.4, 11.5, 12.10, 13.2, 13.6, 15.14
Throw 3.31, 5.3
see pitch, span

XXXV
TIG 3.29
Time constant 3.51, 3.71, 11.1, 11.4, 14.6, 14.9, 14.19, 14.21, 15.15-15.16, 15.19-15.20,
15.22
Time-ratio control 2.15
Tolerance 10.14, 10.23, 13.6, 13.9, 13.22, 14.29, 16.10
Tooth 1.1, 2.3, 3.33-3.S5, 4.27,4.28, 5.5, 5.57, 5.58, 6.21, 8.3, 8.7, 8.93.19, 9.3, 9.5-98.
10.28, 12.5, 12.12, 12.16, 12.23, 12.26, 13.3, 13.7, 13.18, 13.22, 13.23
Tooth span 3.33
Tooth rips 4.28, 13.7
Tooth width 8.11, 12.5, 12,16, 13.3, 13.22
Torque 1.1, 1.2, 1.8, 1.9, 1.11-1.19, 2.1-2.S, 2.11, 2.13-2.15, 2.30, 2.38, 2.43-2.47, 2.49.
3.2-3 6, 3.8, 3.10, 3.13, 3.17-3.20,3.26, 3.34, 3.36, 3.46, 3.47, 3.50, 3.51,
5.71, 3.74, 3.76, 4.1, 4.4, 4.9, 4.264.28, 4.33, 5.5, 5.8, 5.9, 5.12,
5.17-5.21.5.24-5.26,5.29-5.33,5.36-5.39,5.41,5.42,5.57,5.62,5.63, 6.1,
6.3, 6.4, 6.8, 6.10-6.12, 6.22, 6.36, 6.38, 6.39, 6.42-6.44, 6.46, 6.47, 7.1,
7.3, 7.4, 7.6, 7.7, 7.12, 7.14-7.16, 8.1. 8.10, 8.19, 10.2, 10.22, 10.24,
10.28, 10.3010.35, 11.1, 11.2, 11.5-11.12. 12.2-12.4, 12.12, 12.14,
12.19-12.21, 12.28, 13.1. 13.2. 13.7, 13.11, 13.13, 13.14, 13.16, 13.18,
13.19, 13.22, 13.26, 14.1, 14.5. 14.7-14.10, 14.12, 14.14, 14.18, 14.31,
14.41, 15.1, 15.9, 15.17, 15.18, 16.7
linearity 11.2
Torque constant 1.2, 1.13, 1.17, 3.19, 3.51, 4.9, 5.12, 5.36, 5.38, 5.39, 5.42, 6.42, 7.1,
7.3, 7.4, 7.6, 7.7, 7.12, 7.14, 13.11, 14.5
see krpkg
Torque linearity 3.74, 5.38, 11.2, 11.8-11.10
Torque per unit volume 16.7
Torque ripple 1.19, 2.13, 2.46, 2.47, 2.49,3.5, 3.6, 3.19, 3.50,4.264.28, 4.33. 5.19, 5.20.
5.25, 5.26, 5.29, 5.30, 5.62, 5.63, 11.2, 11.10. 11.11, 14.1
Torque/speed
see speed/torque
Torsional 14.9
Toraional resonance 14.9
Torsional stiffness 14.9
Traction 1.12
Transducer 1.10,1.19,1.22-1.25, 2.22,11.6,11.9-11.11,14.13,14.14, 14.18,14.32,14.34
Transfer function 14.5, 14.6, 14.8, 14.12, 14.16, 14.19-14.22, 14.24-14.26, 14.28, 14.29,
14.36, 14.40, 14.43
Transformer 3.15, 8.4, 16.16
Transformer EMF 3.15, 8.4
Transient heating 15.20
Transistor 1.10, 2.16,2.18-2,24, 2.26, 2.27, 2.30, 2.32-2.38, 2.43,2.47,5.5,5.18, 7.2,10.7,
10.9, 10.10, 10.12-10.14. 10.18, 10.22-10.24, 10.26-10.28, 10.30, 13,13
TRV 12.2-12.5
Tuning 14.1, 14.31-14.33, 14.43
Turns 1.7,2.23,2.32, 2.33,3.1, 3.4, 3.10, 3.12,3.15,3.18-3.20, 3.23,3.26, 3.27,3.31-3.33,
3.35, 3.36, 3.38, 3.43, 3.45, 3.47, 3.48, 3.51, 3.70, 3.71, 4.1, 4.194.22,
5.1, 5.3, 5,6, 5.8, 5.9, 5.21, 5.30, 5.35, 5.38, 5.39, 5.42, 5.44, 5.46, 5.47,
5.49-5.51, 5.59, 5.63, 6.4, 6.0.8, 6.13, 6.14, 6.17, 6.206.23, 6.25, 6.27,
6.28, 7.11, 7.14, 8.4, 8.10, 8.17, 10.9, 10.14, 10.18, 10.22, 10.24, 11.2,

xxxvi
I n d ex
11.7,12.3,12.7,13.2,13.6,13.7,13.11,13.12,13.15,13.16,13.22,13.25.
13.26, 16.8
Turns in series per phase 3.27, 4.21, 5.35, 5.S8, 5.39, 5.46, 6.7, 6,13, 6.14, 6.27. 6.28.
7.11, 12.3
Turns per coil 3.23, 3.70, 5.21, 5.59, 13.2, 13.11, 13.25, 13.26
Two-axis theory 2.46, 6.9
Two-phase 1.19, 2.17, 2.26, 2.27, 3.4, 3.27, 4.19,5.32, 5.33, 5.43, 6.31, 6.32, 7.1, 7.7-7.9,
7.13, 12.22, 13.14
Two-phase-on 4.19, 12.22, 13.14
Type 0 14.22, 14.31
Type 1 14.22
Unipolar 2.47, 3.5, 3.27, 5.32, 10.1, 10.4, 10.23-10.26
Vamish 3.32, 3.29, 3.32, 3.33, 3.35, 3.70, 12.5, 15.24
Vector 1.12, 2.46, 2.47, 2.49, 5.37, 5.43, 5.59, 5.60, 5.66, 6.1, 6.2, 6.11, 6.12, 6.18, 6.19.
12.13, 12.14, 12.19, 12.28, 14.34, 14.45, 15.9
Vector control 1.12, 2.46, 2.47, 2.49, 6.2, 6.11, 14.34
Vector potential 5.43, 5.60, 5.66, 12.13, 12.14, 12.19
Vector rotator 2.46
Velocity 1.9, 1.11, 1.15, 1.19, 1.23, 2.11, 2.47, S.2, 4.30, 5.36, 6.9, 6.11, 6.23, 7.3, 7.4,
7.11, 8.5, 8.11, 11.9, 14.1, 14.5, 14,7, 14.10, 14.13-14.15, 14.18, 14.28,
14.29, 14.32, 14.45, 15.6, 15,7
Voice coil actuator 13.16
Voltage 1.14-1.16, 1.18, 1.25, 2.15-2.18, 2.20, 2.21, 2.23-2.25, 2.30-2.35, 2.38-2.41,
2.43-2.45, 2.47, 3.2, 3.4, 3.14-3.17, 3.20, 3.25, 3.31, 3.32, 3.71, 4.1, 4.3,
4.4,4.18, 5.14, 5.17, 5.32, 5.39-5.41,6.1,6.12, 6.13,6.23,6.3(V6.32, 6.35,
6.36, 6.38,6.40-6.47, 7.S-7.7, 7.10,9.1,9.2,10.1, 10.4.10.8-10.12,10.14,
10.16,10.18,10.19,10.22,10.25-10.33,11.1,11.2,11.5-11.7,11.9-11.14,
13.11, 13.13, 13.15, 13.19, 14.2, 14.5-14.8, 14.10, 14.11, 14.17, 14.35,
15.10, 15.24. 16.10
constant 11.1
Voltage PWM 2.21. 2.30, 10.10
Volls 2.18, 2.20, 2.39, 3.18, 3.19, 5.36, 7.1, 7.4, 7.1], 14.35
Washing machine 2.45
Water brake 11.6
Wattmeter 11.14
Waveform 1.7,1.9,1.11,1.18, 1.21, 2.15, 2.23, 2.33-2.35, 2.38, 2.43, 3.6, 3.19, 3.45-3.47,
3.49, S.50, S.74, 4.28, 4.30, 5.3-5.5, 5.8, 5.9. 5.13, 5.14, 5.17-5.20, 5.26,
5.30, 5.32, 5.40, 5.41, 5.65, 7.2, 7.4-7.7, 7.12, 7.15, 7.16,8.1^8 5,8.8-8.10,
8.12-8.19, 9.2, 9.3, 9.5-9 9, 10.1, 10.2, 10.4. 10 28-10.33, 11.5, 11.14,
12.9, 12.14, 12,27, 13.11, 14.2
Wind-up 14.31, 14.32, 14.41
Winder 13.18
Winding 1.1. 1.4, 1.14, 1.16, 1.17, 2.1, 2.6, 2.8, 2.10, 2.12, 2.13, 2.15, 2.27, 2.43, 2.45,
3.1-S.3, 3.5, 5.10, 5.12, 3.14-3.16, 5.18-5.20, 3.24, 3.26, S.28-S.32,
3.34-3.36,3.39-3.51,3.70,3.71,3.734.1-4.2,4.7-4.8,4.10,4.12,4.1&4.21,
4.26.4.32.5.1-5.3,5.5, 5.6,5.9-5.14,5.17-5.21, 5.24-5.26,5.28-5.32,5.35,
5.38, 5.39, 5.41-5.44, 5.47-5.51, 5.54.5.57-5.66,6.1-6.13,6.14, 6.16-6.23,

xxxvii
6.25, 6.27, 6.28, 6.47, 7.5, 7.11-7.16,8.3.8.9, 8.11), 8.17,11.4,11.7,11.8.
12.5. 12.7-12.11, 12.18, 12.19-12.21, 12.23-12.24, 13.2, 13.4, 13.6-13.8,
13.11-13.13, 13.18. 13.22, 13.23, 13.25, 13.26, 15.1-15.2, 15.5, 15.9,
15.11, 15.12. 15.14, 15.15, 15.17, 15.24, 16.7, 16.17
delta 5.21
fly 5.21
integral slot 3,12, 3.34
single needle 5.21
Winding factor 3.34, 5.28-5.30, 5.49, 6.17, 6.19, 6.21, 6.22, 7.12, 7.13
Winding pitch 3.12, 3,14, 3.35, 5.14, 5.58, 13.6-13.8
Winding techniques 2.12, 5.29
Wire 2.30, 2.38,3.1,3.28, 3.30-3.33,3.70,5.11, 5.21, 5.31,11.14,12.6,12.7,13.12.13.13,
13.16, 13.18, 13.25, 13.26, 15.8, 15.12, 16.1, 16.16, 16.17
Wire diameter 3.70, 13.12, 13.13, 13.26
Wire size 3.1, 3.70, 12.7
Wire-EDM 3.28
Word 14.34, 14.35
Work hardening 16.15
Work-hardening 16,15
Wyc-della switching
see star-delta switching
Yoke 2.2, 2.3, 2.6, 2.11, 2.13, 3.6, 3.25, 3.34, 4.2, 4.16, 4.17, 5.31, 5.61, 5.62, 8.10, 8.16,
8.18, 8.19, 9.2,9.3,9.5-9.7,12.4.12.12,12.23.12.26,13.2,13.3,13.9,13.20,13.21.13.23,
15.24
Z-transform 14.1, 14.38
Zero-order hold 14.35, 14.36
Zero^olt loop 2-21, 2.23, 2.33
ZOH 14.85

xxxviii

S-ar putea să vă placă și