Sei sulla pagina 1di 298

Design of an Interior Permanent

Magnet Machine with Concentrated


Windings for Field Weakening
Applications
By
Lester Chong
A thesis submitted to
THE UNIVERSITY OF NEW SOUTH WALES
in partial fulfilment of the
requirements for the degree of
Doctor of Philosophy
(Electrical Engineering)
August, 2011
i
ACKNOWLEDGEMENTS
I am extremely grateful to my two supervisors Professor Faz Rahman and Dr Rukmi Duttafor
all their time, guidance and invaluable advice given to me over the duration of my PhD. I
thankfully acknowledge the inputs from my examinersand Associate Professor J ohn Fletcher. I
wouldalsolike to thank the Australian Government and the University of New South Wales for
sponsoring my studies.
I am very grateful to the wonderful staff and at School of Electrical Engineering and
Telecommunications, especially Dr Baburaj Karnayil, Richard Tuck andGamini Liyadipitiya. I
am also grateful for the advice given by Subash and Seetha from the school of mechanical
engineering. I would liketo thank Dr Howard Lovatt and Colin Bilson at CSIRO for their help
and advice.
A special thank you to my family, especially my mother J enny Ong who has selflessly worked
so hard to raise and support us; my fianc J anice and the Liao family for their support from the
start to the final stages of my PhD. Last but not least I would like to thank mywonderful friends
and colleagues for making my time in Sydney so special and unforgettable.
ii
ABSTRACT
This thesis presents the design of an interior permanent magnet (IPM) machine with
concentrated windings (CW) for field-weakening applications. The initial phase of this
work involved a feasibility study and comparison with the CW surface permanent
magnet machine. Subsequently a CW-IPM machine was designed and constructed with
the aim of achieving a wide constant power speed range (CPSR). Lastly, based on the
constructed design, scalability and efficiency studies were performed.
The work done in this thesis has led to the successful construction of a prototype
machine achieving a very wide 7.2:1 CPSR. At the time of writing, there is no available
literature that clams such a wide CPSR in a concentrated wound permanent magnet
machine. Distributed winding machines capable of achieving such a CPSR have
complex rotors and hence manufacturing issues. The proposed design was subjected to
the same size constraint as two previously constructed 550W distributed winding IPM
machines. With this constraint, the advantage of shorter end winding length was
exploited and the effective length of the machine was increased. This resulted in a
significant increase in output power to 800W throughout the CPSR. A detailed study on
losses performed in this work showed that despite the increased harmonic content
generated by CW, frequency related losses can be minimized through design methods,
and over an 80% efficiency can be achieved throughout the CPSR. Mechanical stress
analysis of the rotor indicated that iron bridges were required between poles to prevent
excessive stress inflicted on the inter-pole link sections during high speed operation.
Based on confidence gained from the experimental verification of the CW-IPM machine
design, a scalability study was performed and designs up to 30kW were proposed. A
iii
study on efficiency optimization was also carried out and the prototype machine was
redesigned to produce up to 93% efficiency.
The work done in this thesis has setup a strong basis for future work on CW-IPM
machines for automotive traction drives and has also proven that this machine type is
suitable for high performance industrial applications requiring high efficiency over a
very wide CPSR.
iv
TABLE OF CONTENTS
ACKNOWLEDGEMENTS .................................................................................. i
ABSTRACT ............................................................................................................ ii
TABLE OF CONTENTS ...................................................................................... iv
LIST OF FIGURES ............................................................................................... x
LIST OF TABLES ................................................................................................. xix
NOMENCLATURE .............................................................................................. xx
CHAPTER 1 ........................................................................................................... 1
1. INTRODUCTION ........................................................................................... 1
1.1 GENERAL BACKGROUND ............................................................................... 1
1.2 LITERATURE REVIEW .................................................................................... 9
1.2.1 IPM Machine Technology .......................................................................... 9
1.2.2 Concentrated Non-overlapping Windings in AC Machines ...................... 16
1.3 SCOPE AND ORGANISATION OF THESIS .......................................................... 25
CHAPTER 2 ........................................................................................................... 28
2. NUMERICAL METHODS FOR THE ANALYSIS AND PREDICTION OF
MACHINE PARAMETERS ............................................................................. 28
2.1 INTRODUCTION .............................................................................................. 28
2.2 THE FINITE ELEMENT METHOD............................................................................. 30
2.2.1 Brief Background of Finite Element Analysis ........................................... 31
2.2.2 Mathematical Formulations for the Physical Model ................................. 31
2.2.3 Discretisation of the Study Domain ........................................................... 34
2.2.4 Defining Boundary Conditions .................................................................. 37
2.2.5 Galerkins Method for Deriving Finite Element Equations ...................... 39
2.2.6 Solving Finite Element Equations with Newton Raphson Method ............ 41
2.2.7 Process of a Time Stepping Finite Element Model ...................................... 43
v
2.3 FINITE ELEMENT METHOD FOR DETERMINING MACHINE PARAMETERS AND
PERFORMANCE CHARACTERISTICS.................................................................. 45
2.3.1 Construction of Geometry and Assignment of Mesh ................................. 45
2.3.2 Defining Material Properties ..................................................................... 47
2.3.3 Coupling of Electrical Circuits .................................................................. 48
2.3.4 Performance Calculation .......................................................................... 49
2.4 CONCLUSION ................................................................................................ 53
CHAPTER 3 ........................................................................................................... 54
3. INVESTIGATION OF THE CONCENTRATED WINDING IPM FOR WIDE
FIELD WEAKENING APPLICATIONS ..................................................... 54
3.1 INTRODUCTION ............................................................................................. 54
3.2 CHOICE OF SLOT AND POLE COMBINATION .................................................. 55
3.2.1 Winding Factor and EMF ........................................................................... 57
3.2.2 Cogging Torque ......................................................................................... 62
3.3 PERFORMANCE IN COMPARISON TO DISTRIBUTED WINDING IPMMACHINE 65
3.3.1 Airgap Flux Harmonics ............................................................................. 65
3.3.2 Saliency Ratio and Constant Power Capability ........................................ 68
3.3.3 End Winding Length .................................................................................. 73
3.3.4 Slot-fill Factor ........................................................................................... 74
3.4 COMPARING THE IPM AND SPM MACHINES WITH CONCENTRATED
WINDINGS ..................................................................................................... 75
3.4.1 Airgap Flux Produced by the Magnets ...................................................... 76
3.4.2 Constant Power Speed Range .................................................................... 79
3.5 CONCLUSION ................................................................................................. 81
CHAPTER 4 ........................................................................................................... 82
4. DESIGN OF A 1KW IPM MACHINE WITH CONCNETRATED WINDINGS
FOR ACHIEVING A VERY WIDE CPSR ................................................... 82
4.1 INTRODUCTION ............................................................................................. 82
4.2 CONDITIONS FOR MAXIMISING THE CPSR .................................................... 83
4.3 CHOICE OF WINDING ARRANGEMENT ........................................................... 88
4.4 MATERIAL CONSIDERATIONS ........................................................................ 90
4.4.1 Permanent Magnet Material ..................................................................... 90
vi
4.4.2 Core Material ............................................................................................ 92
4.4.3 Stator Coil and Insulating Material .......................................................... 94
4.5 OPTIMISATION OF MACHINE GEOMETRY .................................................... 96
4.5.1 Stack Length ............................................................................................ 96
4.5.2 Rotor Outer Diameter and Airgap Length .............................................. 99
4.5.3 Stator Geometry ...................................................................................... 100
4.6 ROTOR DESIGN AND STRUCTURAL CONSIDERATIONS ................................. 104
4.6.1 Structural Considerations ....................................................................... 106
4.7 FINAL MANUFACTURED DESIGN ................................................................. 112
4.7.1 Back EMF from the Finite Element Model ............................................. 114
4.7.2 Cogging Torque from the Finite Element Model .................................... 116
4.7.3 Inductance and Saliency Ratio from the Finite Element Model ............. 116
4.7.4 Torque Performance from the Finite Element Model ............................. 118
4.8 CONCLUSION ............................................................................................... 119
CHAPTER 5 ........................................................................................................ 120
5. ANALYSIS OF LOSSES IN AN IPM MACHINE WITH CONCENTRATED
WINDINGS FOR FIELD WEAKENING APPLICATIONS ...................... 120
5.1 INTRODUCTION ............................................................................................ 120
5.2 MMF HARMONICS AND LOSSES IN MACHINES WITH CONCENTRATED
WINDINGS ................................................................................................... 121
5.3 CORE LOSS .................................................................................................. 127
5.3.1 Comparison of Steel Grades ................................................................... 129
5.3.2 Core Loss of the Final Design ................................................................ 131
5.4 MAGNET LOSS ............................................................................................. 132
5.4.1 Comparison SPM and IPM Magnet Losses ............................................ 134
5.4.2 Effects of Magnet Segmentation .............................................................. 136
5.4.3 Magnet Loss of the Final Design ............................................................ 139
5.5 STATOR WINDING LOSS .............................................................................. 140
5.6 MECHANICAL LOSSES ................................................................................. 143
5.7 FIELD-WEAKENING PERFORMANCE WITH THE INCLUSION OF LOSSES FROM THE
FINITE ELEMENT MODEL ............................................................................. 147
5.8 CONCLUSION .............................................................................................. 148
vii
CHAPTER 6 ........................................................................................................ 149
6. VECTOR CONTROL OF THE IPM MACHINE WITH CONCENTRATED
WINDINGS .................................................................................................... 149
6.1 INTRODUCTION ............................................................................................ 149
6.2 CONTROL METHODOLOGY .......................................................................... 152
6.2.1 Basic Equations Describing the PM Machine ........................................ 153
6.2.2 Variation of Current Phase Angle .......................................................... 154
6.2.3 Current and Voltage Limits ..................................................................... 157
6.2.4 Maximum Torque per unit Current and Field-weakening Trajectories .. 161
6.3 CONTROLLER ARCHITECTURE ..................................................................... 164
6.3.1 Three phase Inversion Technique ........................................................... 165
CHAPTER 7 ........................................................................................................ 170
7. CONSTRUCTION AND PERFORMANCE ANALYSIS OF THE
CONCENTRATED WOUND IPM MACHINE PROTOTYPE ............... 170
7.1 INTRODUCTION ............................................................................................ 170
7.2 CONSTRUCTION PROCESS ............................................................................ 171
7.2.1 Manufacturing Duration ......................................................................... 171
7.2.2 Rotor Assembly ....................................................................................... 173
7.2.3 Stator Core Assembly and Stator Winding ............................................. 175
7.3 OPEN CIRCUIT PARAMETERS ....................................................................... 178
7.3.1 Back EMF ............................................................................................... 178
7.3.2 Cogging Torque ...................................................................................... 179
7.3.3 Inductance and Saliency Ratio ................................................................ 182
7.4 STEADY STATE ANALYSIS ........................................................................... 185
7.4.1 Torque and Power Characteristics ......................................................... 185
7.4.2 Steady State Voltage and Current Characteristics ................................. 189
7.5 TRANSIENT RESPONSE UNDER MTPA OPERATION ..................................... 192
7.5.1 Transient Voltage and Current Characteristics ...................................... 192
7.5.2 Torque Transients ................................................................................... 194
7.6 PERFORMANCE COMPARED TO DISTRIBUTED WINDING IPMMACHINES ..... 196
7.6.1 Power and Torque versus Frequency Comparison ................................. 198
7.6.2 Cogging Torque Comparison .................................................................. 199
viii
7.6.3 Efficiency Comparison ............................................................................ 201
7.6.4 Magnet Volume Comparison .................................................................. 202
7.7 CONCLUSION ............................................................................................... 203
CHAPTER 8 ........................................................................................................ 204
8. EFFICIENCY OPTIMISATION AND SCALABILITY OF THE
CONCENTRATED WINDING IPM MACHINE ........................................ 204
8.1 INTRODUCTION ............................................................................................ 204
8.2 EFFICIENCY OPTIMISATION ......................................................................... 205
8.2.1 Proposed Designs for Efficiency Optimisation ....................................... 206
8.3 SCALABILITY OF THE CONCENTRATED WINDING IPMMACHINE ................ 210
8.3.1 Airgap Length Variation ......................................................................... 210
8.3.2 Magnet Strength and Armature Current Variation ................................. 211
8.3.3 Effect of Scaling the Machine Size .......................................................... 214
8.4 CONCLUSION ............................................................................................... 220
CHAPTER 9 ........................................................................................................ 221
9. CONCLUSION AND SUGGESTION FOR FUTURE WORK ................ 221
9.1 CONCLUSION OF THIS THESIS ...................................................................... 221
9.2 SUGGESTIONS FOR FUTURE WORK ............................................................. 225
REFERENCES .................................................................................................... 227
APPENDIX A ...................................................................................................... 243
A. AC STANDSTILL TEST APPLIED TO THE FINITE ELEMENT MODEL OF THE
SEGMENTED IPMMACHINE .......................................................................... 243
A.1 Results of AC Standstill Test Implemented on the Segmented IPM Machine 243
APPENDIX B ....................................................................................................... 245
B. SALIENCY RATIO OPTIMISATION .................................................................. 245
B.1 Optimisation of Saliency Ratio by Variation of Rotor Magnet Shape ........... 245
ix
APPENDIX C ...................................................................................................... 249
C. INDUCTANCE WAVEFORMS AND SALIENCY RATIO FOR VARIOUS SLOT/POLE
COMBINATION AND FOR DOUBLE-LAYER WINDINGS .................................. 249
C.1 Inductance Waveform and Saliency Ratio Comparison of Various Slot/pole
Combinations ............................................................................................. 249
C.2 Inductance Waveform and Saliency Ratio Comparison with Double-Layer
Windings .................................................................................................... 251
APPENDIX D ...................................................................................................... 253
D. THERMAL MODEL ......................................................................................... 253
D.1 Thermal Model Approximating Temperature at Various Parts of the
Machine ...................................................................................................... 253
APPENDIX E ....................................................................................................... 255
E. FINAL MACHINE DRAWINGS ........................................................................ 255
E.1 ABB Casing used (with Original Induction Motor) ..................................... 255
E.2 Stator of the CW-IPM Prototype ................................................................. 256
E.3 Rotor of the CW-IPM Prototype .................................................................. 257
E.4 Shaft of the CW-IPM Prototype ................................................................... 259
E.5 Key (shaft) of the CW-IPM Prototype ......................................................... 260
E.6 End-plates of the CW-IPM Prototype .......................................................... 261
APPENDIX F ....................................................................................................... 262
F. EXPERIMENTAL SETUP ................................................................................. 262
F.1 The Experimental Setup ............................................................................... 262
F.2 Control Algorithm ......................................................................................... 264
F.3 3-phase IGBT Inverter .................................................................................. 267
F.4 Kollmorgen PM Machine Specifications ...................................................... 270
APPENDIX G ...................................................................................................... 271
G. PUBLICATION LIST ....................................................................................... 271
x
LIST OF FIGURES
Fig. 1.1 Classification of AC machine typesused for traction
applications
2
Fig. 1.2 Various IPM motor geometries 4
Fig. 1.3 Various stator winding layouts 6
Fig. 1.4 Typical PMSM drive system block diagram 7
Fig. 1.5 Ideal field-weakening characteristics of a drive system 8
Fig. 2.1 Various methods to solve Maxwell equations and predict
machine performance
28
Fig. 2.2 Typical finite elements 35
Fig. 2.3 Two dimensional triangular element 36
Fig. 2.4 Dirichelet boundaries 38
Fig. 2.5 Machine with quarter cyclic-symmetry 39
Fig. 2.6 Newton Raphson method 43
Fig. 2.7 Flow chart for the time-stepping finite element process 44
Fig. 2.8 Prototype machine geometry showing regions 45
Fig. 2.9 DW-IPM showing the repletion of phase coils every quarter 46
Fig. 2.10 Mesh structure of the CW-IPM machine 47
Fig. 2.11 B-H Curve for hard and soft magnetic materials 48
Fig. 2.12 Representation of core material characteristics in terms of an
analytic solution
49
Fig.2.13 Representation of permanent magnet material demagnetisation
characteristics
49
Fig. 2.14 Three-phase star conneted circuit with current excitation 50
Fig. 2.15 Meshing of a three-layer airgap 52
Fig. 3.1 14-pole IPM machine with various slot and pole combinations 56
Fig. 3.2 EMF phasor diagram 59
xi
Fig. 3.3 3-phase EMF waveforms and corresponding frequency
spectrum over 1 electrical cycle for 14-pole IPM machine
with different slot and pole combinations
61
Fig. 3.4 Cogging torque waveforms for various slot and pole
combination
63
Fig. 3.5 Flux waveform and corresponding frequency spectrum of a
14-pole, double-layer, DW machine model
66
Fig. 3.6 Flux waveform due to armature reaction for single- and
double-layer CW
67
Fig. 3.7 14-pole DW IPM with flux being channeled to the d- and q-
axis
68
Fig. 3.8 14-pole, 18-slot, CW-IPM flux plot showing no obvious d or
q-axis flux paths
69
Fig. 3.9 Inductance waveform measured from UNSW Segmented IPM
machine
71
Fig. 3.10 Inductance waveform of an 18-slot, 14-pole CW-IPM
machine
71
Fig. 3.11 d- and q-axis inductance comparison 72
Fig. 3.12 Estimated reduction in end winding length 73
Fig. 3.13 Advanced winding methods to achieve a high saliency ratio 74
Fig. 3.14 Performance comparison of different magnet shapes 75
Fig. 3.15 Flux density plot of the v-shaped IPM model showing
saturation regions
77
Fig. 3.16 Airgap flux produced by the different rotor configurations 78
Fig. 3.17 Peak power and CPSR comparison between three rotor types 79
Fig. 4.1 Rotor types used for increasing saliency ratio 83
Fig. 4.2 Contour plot showing the variation of magnet remanent flux
density versus rated current
87
Fig. 4.3 Permanent magnet demagnetisation curve 90
Fig. 4.4 Magnetisation curve of 35RM300 93
Fig. 4.5 Core loss curve of 35RM300 94
xii
Fig. 4.6 Resistance and copper loss per phase as temperature increases 95
Fig. 4.7 ABB casing with inserted 18-slot stator core 96
Fig. 4.8 End winding length comparison between UNSW DW stator
with windings done on a plastic model
97
Fig. 4.9 Hand winding methods 98
Fig. 4.10 Airgap length variation with output power and CPSR 99
Fig. 4.11 Key parameters defining the stator geometry 100
Fig. 4.12 Flux density plots showing peak flux density in the yoke and
tooth
102
Fig. 4.13 Plastic stator made with three different slot opening widths 103
Fig. 4.14 IPM rotors showing inter-pole link sections 104
Fig. 4.15 V-angle variation 105
Fig. 4.16 Normalised power versus speed characteristics with variation
of v-angle
105
Fig. 4.17 Various types of SPM rotors 106
Fig. 4.18 IPM rotor with buried single-piece/pole magnets 107
Fig. 4.19 Modelled solenoid-magnet model 108
Fig. 4.20 Sections for centrifugal force calculation 109
Fig. 4.21 Model showing outward normal pressure on each section of
the rotor
109
Fig. 4.22 Mechanical stress analysis of rotor 110
Fig. 4.23 Final 18-slot, double-layer winding layout 112
Fig. 4.24 Three-phase induced line to neutral back EMF voltage from
the FE model (at 50Hz)
114
Fig. 4.25 Induced line to line voltage versus speed 114
Fig. 4.26 Comparison of EMF waveform between the CW-IPM and
DW-IPM
115
Fig. 4.27 Comparison of EMF waveform between the CW-IPM and
DW-IPM frequency spectrum
115
xiii
Fig. 4.28 Cogging torque of final design comapred to an equivaent
integral-slot DW model
116
Fig. 4.29 Self- and mutual-inductance waveform of final model with
3A
rms
current excitation
117
Fig. 4.30 Torque ripple of final model at base speed 118
Fig. 4.31 CPSR of final design with base frequency of 50Hz 118
Fig. 5.1 Comparison of flux distribution in a CW and DW machine 121
Fig. 5.2 Ref. to fig. 3.5 122
Fig. 5.3 Ref. to fig. 3.6 122
Fig. 5.4 Contribution of eddy current and hysteresis loss 124
Fig. 5.5 Ref. to fig. 4.5 128
Fig. 5.6 Annular steel model to determine eddy current and hysteresis
loss constants
128
Fig. 5.7 Core loss comparison at 50 and 500Hz with different steel
grades
130
Fig. 5.8 Core loss with chosen steel grade at various frequencies 131
Fig. 5.9 Extrapolated values of measured hysteresis and eddy current
loss for sintered neodymium magnets at 50Hz
133
Fig. 5.10 3D model of a single-pole and single-phase excitation v-
shaped IPM
134
Fig. 5.11 3D model of a single-pole and single-phase excitation SPM 134
Fig. 5.13 Comparison of magnet losses between IPM and SPM machine 135
Fig. 5.14 Variation of magnet eddy current loss with number of magnet
segments due to slot and carrier harmonics Comparison
between IPM, inset and SPM rotors
137
Fig. 5.15 Circulating eddy currents in a non-segmented a) SPM magnet
pole, b) V-shaped IPM magnet pole
137
Fig. 5.16 Circulating eddy currents in a segmented magnet poles 138
Fig. 5.17 Magnet losses in an SPM machine with variation of segment
number
138
xiv
Fig. 5.18 Magnet losses in an IPM machine with variation of segment
number
139
Fig. 5.19 Magnet losses versus frequency in the final design with
sintered NdFeB magnets
140
Fig. 5.20 Inner-diameter view of stator teeth showing estimated
winding length
142
Fig. 5.21 Drawing of rotor indicating key components to calculate
bearing friction loss
143
Fig. 5.22 Power loss due to bearing friction at various speeds 144
Fig. 5.23 Power loss due to windage at various speeds 146
Fig. 5.24 Modelled power versus speed performance with inclusion of
losses
147
Fig. 6.1 2-pole IPM machine showing d-q axis reference frames 150
Fig. 6.2 Space vector dq-axis phasor diagram 151
Fig. 6.3 Calculated torque comprising of alignment and reluctance
torque
154
Fig. 6.4 Current phasor under MTPA and Field-weakening operation 155
Fig. 6.5 Back EMF, induced current and induced voltage waveforms
under MTPA and maximum field-weakening operation
156
Fig. 6.6 Circle diagram for IPM and SPM machine showing current
and voltage limits of the system
157
Fig 6.7 Classificationof machine type by characteristic current 160
Fig. 6.8 i
d
and i
q
field-weakening current trajectory for the CW-IPM
model
161
Fig. 6.9 Comparison of d-axis current trajectories by Morimotos
equations and through repetitive testings
163
Fig. 6.10 Vector control system block diagram 164
Fig. 6.11 Rectifier Inverter for producing three-phase outputs to the
machine
165
Fig. 6.12 Switching vectors of the space vector modulation method 166
Fig. 6.13 Switching pattern for sector 1 168
xv
Fig. 7.1 Duration breakdown and steps of construction process 172
Fig. 7.2 Laser-cut rotor lamination 173
Fig. 7.3 Complete rotor stack with magnets inserted 174
Fig. 7.4 Completed dynamically-balanced rotor 174
Fig. 7.5 Completed stator stack 175
Fig. 7.6 Measurement of end winding length in UNSW CW-IPM
stator
176
Fig. 7.7 Comparison of UNSW CW-IPM machine assembly and DW
S-IPM machine
177
Fig. 7.8 Measured line to line and line to neutral back EMF
waveforms compared against modelled values
178
Fig. 7.9 Measured line to line back EMF versus speed compared
against modelledvalues
179
Fig. 7.10 Cogging torque measurement setup 179
Fig. 7.11 Measured cogging torque points compared against results
obtained from FE model
180
Fig. 7.12 Curve fitted cogging torque waveform 181
Fig. 7.13 Ref. to fig. 4.29 182
Fig. 7.14 Measured self and mutual-inductance waveform from the
prototype with 3A
rms
current excitation
183
Fig. 7.15 Variation of dq-axis inductances with current 184
Fig. 7.16 Experimental setup to measure back EMF and torque/power
versus speed performance
185
Fig. 7.17 Measured dq-axis current points under field weakening
operation
186
Fig. 7.18 Measured line to line voltage versus speed 187
Fig. 7.19 Measured torque versus speed characteristics of the CW-IPM
machine prototype
187
Fig. 7.20 Measured efficiency versus speed 188
Fig. 7.21 Steady-state speed waveforms 189
xvi
Fig. 7.22 Steady state i
d
current waveforms 190
Fig. 7.23 Steady state i
q
current waveforms 190
Fig. 7.24 Steady state current and voltage inputs to the machine in abc
reference frame
191
Fig. 7.25 Speed step from standstill to base speed 192
Fig. 7.26 i
d
current waveforms with speed step from standstill to base
speed
192
Fig. 7.27 i
q
current waveforms with speed step from standstill to base
speed
193
Fig. 7.28 Current and voltage inputs to the machine in abc reference
frame with step change in speed
193
Fig. 7.29 Torque transient when CW-IPM machine accelerates from
standstill to base speed at full load
194
Fig. 7.30 Measured torque ripple at steady state 195
Fig. 7.31 Comparison of three UNSW IPM machines 197
Fig. 7.32 CPSR comparison between the three UNSW IPM machines 198
Fig. 7.33 Normalised output torque comparison between the three
UNSW IPM machines
199
Fig. 7.34 Cogging torque comparison between the three UNSW IPM
machines
200
Fig. 7.35 Cogging torque as a percentage of output torque at base speed
comparison between the three UNSW IPM machines
200
Fig. 7.36 Efficiency comparison between the CW-IPM and S-IPM
machine up to 200Hz
201
Fig. 7.37 Comparison of magnet volume per kW the three UNSW IPM
machines
202
Fig. 8.1 Axial length comparison between two different slot-fill
methods
206
Fig. 8.2 Designs used for efficiency optimization indicating outer
dimensions
207
xvii
Fig. 8.3 Efficiency andpower versus speed performance of the two
efficiency optimised model
208
Fig. 8.4 Airgap length versus CPSR and input power scalability
study
210
Fig. 8.5 Airgap length versus efficiency and core loss scalability
study
211
Fig. 8.6 Magnet remanent flux density versus input current and input
power scalability study
212
Fig. 8.7 Flux density plot of CW-IPM machine under saturated and
unsaturated conditions
213
Fig. 8.8 Magnet remanent flux density versus total machine losses and
efficiency scalability study
213
Fig. 8.9 Comparison between the three machine sizes used in
scalability study
215
Fig. 8.10 Input power, output power and efficiency versus speed
characteristics of the 5kW design
217
Fig. 8.11 Input power, output power and efficiency versus speed
characteristics of the 30kW design
218
Fig. 8.12 Losses in the three machine sizes as a percentage of total loss
scalability study
219
Fig. A.1 Segmented IPM machine 243
Fig. A.2 AC standstill test on the S-IPM machine using FE analysis
compared to other methods
244
Fig. C.1 Three 14-pole layouts used in the saliency ratio comparison 249
Fig. C.2 Self- and mutual-inductance for the 6-slot, 14-pole, single-
layer model
254
Fig. C.3 Self- and mutual-inductance for the 12-slot, 14-pole, single-
layer model
250
Fig. C.4 Self- and mutual-inductance for the 18-slot, 14-pole, single-
layer model
250
Fig. C.5 Single- and double-layer layout used in the saliency ratio
comparison
251
xviii
Fig. C.6 Self- and mutual-inductance for the 6-slot, 14-pole, double-
layer model
251
Fig. D.1 Approximate model used in thermal analysis 254
Fig. F.1 Complete experimental Setup 262
Fig. F.2 d-space control desk, real-time graphical user interface 263
Fig. F.3 3-phse IGBT inverter (casing off) 267
Fig. F.4 IGBT inverter schematic 268
Fig. F.5 Connections between the IGBT inverter and control boards 269
xix
LIST OF TABLES
Table 3.1 Winding factor and LCM for different S
pp
values 60
Table 3.2 Key specifications of Models with Different Magnet Shapes 76
Table 4.1 Advantages and disadvantages of different magnet types 91
Table 4.2 Power and CPSR with variation of v-angle 105
Table 4.3 Mechanical data for silicon steel grade (J FE-35J N210) 110
Table 4.4 Specifications of the final design 113
Table 5.1 Properties of compared core grades 129
Table 5.2 Comparison of magnet losses between IPM and SPM machine 135
Table 5.3 Suitable conductor sizes and properties 141
Table 5.4 Bearing friction loss on each bearing at various speeds 144
Table 5.5 Reynolds number, torque coefficient and winding loss at
various speeds
146
Table 6.1 Space vector modulation look-up table 167
Table 8.1 Key specifications of the two Optimised CW-IPM Machine
Designs
208
Table 8.2 Key specifications of the three CW-IPM Machine Designs 217
Table C.1 Saliency ratios for various layouts 252
Table D.1 Keymachine parameters used in thermal model 253
Table D.2 Estimated temperature at various parts of the machine 254
Table F.1 Kollmorgen PM machine specifications Loading machine 270
xx
NOMENCLATURE
A
A
g
Airgap area [m
2
]
A
m
Magnet pole area [m
2
]
A
x
Area of slot [m
2
]
A
w
Cross sectional area of a wire [m
2
]
A
Magnetic vector potential
A
z
Z-axis scalar potential
AC Alternating current
B
B
g
Airgap flux density due to the magnets [T]
B
r
Magnet remanent flux density [T]
B
x
Flux density normal to the pole surface [T]
B
Magnetic flux density vector [T]
C
c Circumference [m]
CPSR Constant power speed range
CW Concentrated non-overlapping winding
D
D Diameter [m]
D
os
Stator outer diameter [m]
d-axis Direct axis
d Number of nodal degrees of freedom
DC Direct current
DTC Direct torque control
DW Distributed winding
xxi
E
EMF Electromotive force [V]
e Electromotive force (instantaneous) [V]
F

x
EMF reference phasor of winding element x [V]
F
Electromagnetic field vector
F
f Frequency [Hz]
f
s
Sampling frequency [Hz]
FE Finite element
FOC Field oriented control
FFT Fast Fourier transform
H
H
Magnetic field vector [Hz]
I
|
a
Phase A current [A]
|
h
Phase B current [A]
|
c
Phase C current [A]
I
ch
Characteristic current [A]
I
r
Rated current [A]
IM Induction machine
IMA Integrated motor assist
IPM Interior permanent magnet
ISA Integrated starter-alternator
J
] Electric current density [A/m
2
]
]
a
Equivalent electric current density from armature coils [A/m
2
]
]
m
Equivalent electric current density from magnets [A/m
2
]
xxii
K
k
e
Eddy current loss constant
k
h
Hysteresis loss constant
k
t
Torque constant
k
w
Winding factor
k
p
Pitch factor
k
d
Distribution factor
k
x
Skew factor
L
|
e
Stacklength of machine [m]
|
ag
Airgap length [m]
|
r
Rotor length [m]
L
x

Stator inductance [H]
L
d

d-axis inductance [H]
L
q
q-axis inductance [H]
L
a
Self-inductance of phase A winding [H]
LCM Lowest common multiple
M
M
ah
Mutual inductance between phase A and B winding [H]
MMF Magneto motiveforce [AT]
m Mass of object [kg]
M
Magnetization vector
MTPC Maximum torque per current
N
N
ph
Number of turns per phase [Turns]
N
cu||
Number of turns per coil [Turns]
N
a
Winding function of phase A [Turns]
xxiii
N
h
Winding function of phase B [Turns]
n
x
Rotational speed [RPM]
P
P
m
Magnetic pressure [Pa]
P
eddy
Eddy Current Loss [W]
P
hyx
HysteresisLoss [W]
P
cure
Total core loss [W]
3
x
Magnetic polarisation at saturation [C/m
2
]
p Pole pairs
PM Permanent magnet
PMSM Permanent magnet synchronous machine
PWM Pulse width modulated
Q
q-axis Quadrature axis
R
r Radius [m]
r
g
Airgap radius [m]
R Resistance |O|
3 Reluctance [AT/Wb]
3
g
Reluctance in the airgap [AT/Wb]
R Residual
S
x
d|x
Displacement [m]
Spp Slots per pole per phase
Sff Slot fill factor
x Slots
xxiv
SEG Segmented
SPM Surface permanent magnet
SVM Space vector modulation
T
T Torque [Nm]
t Time [s]
T
e
Electro-magnet/alignment torque [Nm]
T
re|
Reluctance torque [Nm]
T
cug
Cogging torque [Nm]
V
V Voltage [V]
Velocity [m/s]
W
W
|
Weight function
W Virtual work
Speed in radians per second

ref
Speed reference signal
Greek Letters and Symbols
u Temperature coefficient of resistivity [1/k]
A
m
Magnet internal permeance [Wb/A]
q Flux [Wb]
q
ag
Airgap flux [Wb]
q
pm
Permanent magnet flux [Wb]
Flux linkage [Wb]

pm
Permanent magnet flux linkage [Wb]
m
m
Mechanical speed [rad/s]
0
r
Rotor position [degrees]
0
m
Rotor mechanical position [degrees]
xxv
Permeability [H/m]

Permeability of free space [H/m]

r(mag)
Magnet relative permeability [H/m]
p Resistivity |O.m|
Conductivity [S/m]
y Current displacement angle rads
Electric scalar potential
4 Electric scalar potential
I Boundary
Numbers
2D Two-dimensional
3D Three-dimensional
1
CHAPTER 1
1. Introduction
Chapter 1 will give an introduction to the permanent magnet synchronous machine, with
the main focus on the interior permanent magnet (IPM) type for field-weakening
applications. A general overview of the drive system architecture with different modes
of operation will be examined. Two major research areas which are relevant to this
thesis will be reviewed, namely, the IPM machine technology, and the concentrated
non-overlapping windings (CW) used in permanent magnet alternating current (AC)
machines. This chapter will also summarise research in various literatures relevant to
this work. Lastly, this chapter will provide an overall outline of this thesis.
1.1 GENERAL BACKGROUND
Over the years, the application of electric motors has replaced vast numbers of
mechanical rotating devices. From tiny motors used in wristwatches, to very large
motors used for ship propulsion and wind turbines. There are numerous types of electric
motors available for present-day applications, of which the AC type are most commonly
used in high performance applications due to its increased efficiency and excellent
dynamic performance. The classifications of common types of AC motors are shown in
fig. 1.1 [1-4].
Chapter 1: Introduction
2
Fig. 1.1 Classification of AC machine typesused for traction applications
Machine types described in fig. 1.1, are suitable for traction/field-weakening
applications. While work on Halbach machines is still in the research phase, the
induction, SPM, inset, and IPM machine types have already been applied to present day
traction drive systems. Some of the examples include [2]:
Induction machine: GM Chevrolet Silverado hybrid and GMC Sierra hybrid
SPM/Inset PM machine: Honda Insight and Civic hybrids
IPM machine: Toyota Prius and Ford Escapehybrid
Induction, SPM and inset PM machines usually have a lower power rating compared to
the IPM machine and are most commonly applied as an integrated motor assist (IMA)
system, where the main driver of the vehicle is the internal combustion engine while the
electric motor assists. On the other hand, the IPM machine itself produces up to 73kW
of power (for the case of the 3
rd
Generation Prius) and can be driven in full electric
mode, producing zero emissions.
AC
Machines
Synchronous
Induction machine
SPM machine
Inset PM machine
IPM machine
Asynchronous
Halbach machine
Distributed windings
Concentrated windings
Electrically excited
machine
Chapter 1: Introduction
3
This thesis will focus on the IPM machine type, which is generally preferred due to
three main reasons: Firstly, the buried magnets make the rotor structurally stronger,
which make it more capable of withstanding higher speeds. Secondly, the additional
useful reluctance torque, resultingfrom the salient pole structure, thus giving the motor
greater field-weakening capabilities. Additionally this saliency allows sensorless control,
properties which the SPM does not offer. Lastly, the possibility of changing the
geometry of buried magnets in the rotor makes it possible to employ flux concentration,
and provides the possibility of saliency ratio optimisation [3, 5].
With the availability of high energy permanent magnet materials and advanced power
electronics, the fields in which IPM machines can be applied to are rapidly broadening.
They include aerospace, nautical, automobile, rail transportation, medical, generation
and industrial process automation [6-11].
Common magnet geometries include single-piece/pole, rectangular shaped magnet
design (fig. 1.2a), segmented magnet design (fig. 1.2b), v-shaped magnet design (fig.
1.2c), and themulti-barrier design (fig. 1.2d) [12-14].
Chapter 1: Introduction
4
(a) rectangular, single-piece/pole magnets (b) segmented magnets
(c) v-shaped magnets (d) multi-barrier magnets
Fig. 1.2 Various IPM rotor geometries
Each of these designs has its advantages and disadvantages: The single-piece/pole
magnet design, for example, is the easiest to manufacture, but has larger magnet losses
compared to the other designs due to the larger magnet pole surface [15]. The
segmented magnet design has lower magnet losses and better field-weakening capability
but requires more magnet pieces. It alsoresults in decreased magnet flux densitydue to
the leakage flux in the iron bridges [13]. For further explanation of the segmented
machine design, one could refer to [174]. The v-shaped magnet design provides flux
concentration but also requires more magnet pieces compared to the single-piece/pole
Chapter 1: Introduction
5
design. The multi-barrier magnet design creates a very high saliency ratio, but
consequently results in an increased amount of structural stress on the rotor steel [12].
In practice, there is no single rotor that can satisfy all applications. The pros and cons of
each design as well as more specific magnet type and shape have to be altered to meet
desired specifications. As magnets are very brittle, there are also practical limits to the
manufacturability of the magnets.
Before the 21
st
century, the majority of IPM machines were designed with distributed
stator windings (DW). The use of CW was not popular due to a poor torque to
magnetomotive force (MMF) ratio. However in the early 21
st
century, Cros and
Viarouge [16], Magnussen and Sadarangani [17] proved that by an appropriate choice
of slot and pole combination, the winding factor can be significantly increased, thus
increasing output torque. Additionally it was also shown that withappropriateslot and
pole combination, cogging torque canalsobe reduced.
In CW, the opposite polarity of the corresponding phase coil is located in the next slot,
therefore end windings do not overlap. This results in a much shorter end winding
length compared to DW, which is beneficial for applications with space constraints.
Stator windings can either be single- or double-layer. The choice depends on the desired
machine performance characteristics. Single-layer CW creates high self-inductance and
low mutual-inductance which leads to better fault-tollerant capability. On the other hand,
double-layer CW has lower airgap MMF harmonic components, thereby resulting in
smaller torque ripples and lower magnet eddy current losses [18-20]. The winding
layouts for single- and double-layer DW are shown in fig.1.3a and 1.3b, while the
layouts for single- and double-layer CW are shown in fig.1.3c and 1.3d respectively.
Chapter 1: Introduction
6
(a) Single-layer distributed winidngs (b) Double-layer distributed windings
(c) Single-layer concnetrated winidngs (d) Double-layer concentrated widnings
(e) Double-layer modular concentrated widnings
Fig. 1.3 Various stator winding layouts
Chapter 1: Introduction
7
Occasionally, double-layer CW is further classified into subcategories: traditional CW
and modular CW [21-23]. Traditional CW does not have the same phase windings
around consecutive teeth (fig.1.3d), whereas modular CW (fig.1.3e) does. Modular CW
has the advantage of reduced torque ripple; however, it generates increased harmonic
content compared to traditional CW. In this thesis, CW would be only classified as
single- or double-layer and would not be further divided into traditional and modular.
The standard method used to determine the phase coil arrangements is as shown in [16].
Drives for present day applications require fast response and precise control of position,
speed and torque, hence control loops with constantly improving controller algorithms
are required. A commonly used drive system block diagram for a permanent magnet
synchronous machine (PMSM) is shown in fig. 1.4.
Most drive systems in the past made use of rotor position sensors to achieve precise
control. Recently, there has been keen interest in sensorless control schemes, such as
sensorless direct torque control (DTC), which eliminates the need for a physical sensor.
Fig. 1.4 Typical PMSM drive system block diagram
J
Jt
Controller
Power
Converter
Load
Input
Power
PMSM
i
u
i
b
0
m

m
I,
m
PWM
Signal
Chapter 1: Introduction
8
The operation of the drive can be classified into two parts: the constant torque
(maximum torque per unit current (MTPC) region), and the constant power (field-
weakening region). The width of the field-weakening region is also known as the
constant power speed range (CPSR) of the machine. With the current kept at rated
value throughout the entire speed range, voltage increases with speed till its rated value
is reached. This point is the limit of the MTPC region. After which, in order to increase
the speed further, the field of the machine has to be weakened to keep voltages within
operating limits. Field weakening is achieved by splitting the current into two
components namely the d-axis (magnet pole axis) and the q-axis (inter-pole axis)
components. By the decoupling these two currents, each of them could be made to
control the torque and flux independently. In order to maintain a constant power speed
range, the magnitude of negative d-axis current I
d
increased to weaken the magnet field
and maintain constant voltage, inevitably the q-axis current I
q
has to be decreased to
maintain a constant armature current I
a
. The limit to which the fieldof the machine can
be weakening to achieve a wide constant power speed rangedepends on two key factors:
the saliency ratio and the characteristic current. In an ideal case, a drive system would
be able to reach infinite speeds as shown in Fig 1.5.
Chapter 1: Introduction
9
Fig. 1.5 Ideal field-weakening characteristics of a drive system
1.2 LITERATURE REVIEW
DW-IPM machines have been widely used in high performance industrial applications
over decades. CW-IPM machines on the other hand have only more recently found their
way into industrial markets. This section reviews well established work in IPM machine
technology as well as recent research in CW for AC machines.
1.2.1 IPM Machine Technology
The IPM machine has been a popular choice for field-weakening applications. Global
research on the IPM machine dates back to the late 1970s and early 1980s when the first
few papers were published; there is still very significant research interest in this area
today. The popularity of the IPM machine is due to the embedded structure of its
Torque
Input
Voltage
Flux
Speed
m
Armature
Current
Output
Power
Speed
m
Speed
m
Speed
m
t
Speed
m
Constant torque
(MTPC region)
Constant power
(Field-weakening region)
Chapter 1: Introduction
10
magnets, which lowers the risk of demagnetisation, increases the mechanical robustness
of the rotor and provides additional reluctance torque. The small airgap design in most
IPM machines makes it excellent for flux weakening, as the negative armature reaction
can effectively reduce airgap flux. The IPM machine also gives the machine designer
the freedom to vary the magnet pole geometry, thereby broadening the machines area
of application.
Leading studies in IPM machine technology has included patents and several papers
setting the basis for research in this area. Steen [24] filed a patent on synchronous
motors with buried permanent magnets having several geometrical configurations. He
stated that the buried magnets produced additional direct-axis (d-axis) flux in aid of the
flux generated by the inductive copper bars during no-load operation. Honsinger [25]
illustrated a detailed mathematical representation of the IPM machine which included
its magnetic fields and parameters. Rahman et al. [26] presented the equivalent circuit
model to determine the d-axis and quadrature-axis (q-axis) reactance. Consoli and
Renna [27] illustrated a detailed representation of the IPM machine in the rotor
reference frame, and demonstrated an equivalent circuit model to determine iron losses.
Chalmers et al. [28] presented a study of the IPM machine through extensive
experiments with frequency variations. They explained how the q-axis reactance can be
more accurately obtained with the consideration of saturation. One of the first papers to
illustrate the use of finite element (FE) analysis in IPM machine design was by Schiferl
and Lipo [29]. They showed how FE analysis is used to increase the prediction accuracy
of losses at low voltage levels by considering harmonics caused by armature excitation.
Noticing the IPM machines superiority when used in flux weakening applications,
J ahns et al. [30] first addressed the IPM machines characteristics when used as a high-
Chapter 1: Introduction
11
performance variable speed drive. J ahns [31] then performed a novel study on the flux
weakening of the IPM machine, thereby successfully extending its constant power
speed range. Since then, the use of IPM machines for flux weakening operations has
soared. Soonget al. [12, 32] introduced the optimal field-weakeningconditions for IPM
machines using the parameter plane concept. They constructed and compared several
rotor types and proved that the multi-barrier IPM rotor produced the most promising
field-weakening performance. Soong et al. [33, 34] also developed a new axially
laminated IPM motor capable of achieving an extremely wideCPSR. Honda et al. [35,
36] did a study on the effects of various winding types and rotor configuration on the
field-weakening performance of the IPM machine. J olly et al. [37] used genetic
algorithms to determine theoptimal CPSR of an IPM machine confirmingtheir results
with FE analysis.
The IPM machines characteristics were constantly compared to those of other AC
machines. Fratta et al. [38] compared the torque density and flux weakening ability of
the IPM machine with the induction machine, and highlighted that the IPM would have
better electromagnetic performance if mechanical issues were resolved. A
controllability comparison betweenthe IPM and SPM machines under various operating
requirements of the current vector control scheme was done by Morimoto et al. [39].
They concluded that the IPM machine offered better flux-weakening capability. Zhu et
al. [40] compared the iron loss between the IPM and SPM machines. They indicated
that the iron loss of the IPM machine wouldbe lower under open-circuit conditions but
significantly higher in the field-weakeningregion compared to the SPM machine, due to
the increased harmonic content in the armature field. Kyung-Tae et al. [41] compared
the effects of rotor eccentricity on the IPM and SPM rotor, in which theyconcluded that
Chapter 1: Introduction
12
the IPM is more pronetothe effects of rotor eccentricity. J ung Ho et al. [42] studied the
inductance variation of a hybrid synchronous reluctance/IPM motor and found out that
the addition of buried magnets increased the saliency ratio, thus increasing output
torque and power factor.
A substantial amount of research efforts was focused on the control of the IPM machine.
Morimoto et al. [43] very clearly illustrated the control algorithm for an IPM motor with
a high performance current regulator over the entire speed range. Uddin et al. [44]
compared various current controllers for a voltage source-driven IPM drive and
proposed a hybrid type converter to exploit the best characteristics of the different
controllers.
The desire to completely eliminate position sensors and to increase the reliability of
drive systems sparked keen research interest in sensorless control of IPM machines.
Rahman et al. [45] as well as Ogasawara and Akagi [46, 47] were among to first authors
to implement sensorless control on IPM machine. At the same time, Corley and Lorenz
[48] achieved sensorless control over awide speed range includingzero speed. Haque et
al. [49] usedthe high frequency injection method to estimate the initial rotor position in
an attempt to completely eliminate sensors fromIPM drives from start up. Sergeant et al.
[50] studied the effects of rotor geometry on sensorless control. Gyu-Hong et al. [51]
highlighted the effects on machine control performance due to varying L
d
andL
q
under
various load conditions, and proposed a control method with the addition of inductance
estimation.
Cross-coupling between the d- and q-axis fluxes affects machine parameter prediction,
thus studies were made to minimise its effects. Bianchi et al. [52-54] studied the effects
of various rotor structures on cross-coupling and sensorless control, including the high
Chapter 1: Introduction
13
frequency signal injections technique to determine the rotor initial position. They noted
that the multi-flux barrier motor produced the lowest cross-coupling effects and hence
lower saturation. The multi-flux barrier motor with sensorless control was later studied
by Wu at al. [55].
The end of the 20
th
century saw drastic improvements in computational resources and
techniques. This allowed machine designers to efficiently determine parameters and
perform optimisation strategies. Novel attempts to reduce cogging torque with the aid of
two-dimensional (2D) FE analysis by rotor pole or stator tooth shaping was done by
Filho et al. [56]. Fujishima et al. [57] as well as Kano and Matsui [58] adopted the FE
and genetic algorithm search method to derive optimal multi-objective designs.
Yamazaki [59] illustrated a method to calculate IPM machine parameters, including
rotor and stator iron losses, with FE analysis. Efficiency optimisation by geometric
variation was carried out by Sim et al. [60] using finite element models. Dong-Hunet
al. [61] attempted to reduce cogging torque by producing the rotor stack with unequal
lamination in the outer diameters. Ki-Chan et al. [62] studied the effects on machine
parameters and torque performance by varying the shape of rotor link sections. They
showed by FE analysis that small variations of the link sections could significantly
affect cogging torque and saliency ratio. Parsa and Lei [63] studied the effects of torque
ripple and performance characteristics when key machine parameters were varied.
Kioumarsi et al. [64] attempted to reduce torque ripple and increase field-weakening
performance by drilling additional holes in the rotor. Kang et al. [65] also attempted to
reduce cogging torque by rotor surface shaping and by adding notches in the steel. They
showed that the EMF harmonics and cogging torque amplitude could be effectively
reduced. Han et al. [66] attempted to reduce torque ripple by varying the number of
Chapter 1: Introduction
14
slots and number of rotor barriers. They showed that multi-barrier rotors with an odd
number of slots per pole pair resulted in low torque ripple. Sanada et al. [67]
experimented with several designs and proved that the use of asymmetric flux barriers
was beneficial in reducing torque ripple in multi-barrier IPM machines. Fang at al. [68]
showed that torque ripple and cogging torque can be reduced with a double-layer rotor.
Kimat al. [69] studied the effects of geometric variations of magnets in the IPM
machine to reduce torque ripple.
Computational methods also made the calculation of losses more accurate and realisable.
Loss minimisation was also made less costly and more effective. Kawase [70] analysed
permanent magnet eddy current losses in the IPM machine with three-dimensional finite
element (FE) analysis, and showed the effectiveness of reducing eddy current losses by
axially dividing the magnets. Zivotic-Kukolj [71] proposed geometric variations to the
multi-barrier IPM machine to reduce iron losses. Ionel et al. [72] improved the accuracy
of modelling core losses by allowing hysteresis loss to vary with both flux density and
frequency, but leaving the eddy current and excess losses to vary only with flux density.
Yamazaki and Seto [73] studied iron loss of the IPM machine at various operating
speeds, and concluded that at low speeds PWM inverters contribute the most to iron
losses, whereas losses from PM MMF harmonics are more significant during field-
weakening operation. Wang et al. [74] studied the effects of temperature on the torque
performance and machine losses in an IPM machine using FE analysis. Yamazaki and
Kanou [75] as well as Okitsu et al. [76] proposed more efficient yet equally accurate
alternatives to the full three-dimensional (3D) FE method for IPM rotor loss calculation.
Seo et al. [77] studied iron loss on machines with integral and fractional slot DW. They
showed that the with fractional-slot configuration, iron loss in the stator is reduced
Chapter 1: Introduction
15
slightly, but iron loss in the rotor is increased significantly, especially at high speeds.
Yamazaki and Ishigami [78] attempted to reduce iron loss by altering the rotor core
structure by having slits or air-pockets at various parts of the magnet surface. Yamazaki
and Abe [79] further investigated the effects of magnet segmentation in IPM motors,
and concluded that the axial length of each magnet segment should not be more than
twice the skin depth of the eddy currents produced by the dominant harmonics. Han et
al. [80] attempted to minimize eddy current loss in the stator teeth of IPM rotors under
field-weakening conditions and highlighted that double-layer rotor magnets resulted in
lower losses compared to single-layer rotor magnets. Tseng and Wee [81] investigated
various methods to determine core loss in the IPM machine, in which they stated the
relationship between flux and core loss as well as appropriate core loss calculation
methods to use at different stages of the machine design to save resources. Stumberger
et al. [82] studied the iron losses under field-weakening operation and stated that the
rotor iron losses is substantial, despite there being a very small portion of iron is present
above the magnets. Ma et al. [83] proposed a method to increase the accuracy of
calculating iron loss with FE analysis using rotational fields and flux density harmonics.
Barcaro et al. [84] also studied how the design of rotor flux barriers and the amount of
PM material affected the losses in an IPM machine.
Equivalent circuit and analytical models were also continuously improved and used as
an alternative or together with FE models for quicker calculations. Fernandez-Bernal et
al. [85] attemptedto establish parameters for any kind of PM synchronous machine. Lee
et al. [86] improved the equivalent circuit model of the IPM machine taking into
account the effects of saturation. Zhu et al. [87] proposed an analytical model for the
multi-barrier and segmented IPM airgap flux without armature excitation.
Chapter 1: Introduction
16
Structural and thermal issues are closely related to durability, acoustic noise and
efficiency that can be addressed in the design stages of the drive system. Lovelace et al.
[88] addressed mechanical issues regarding the structural stresses inflicted on the multi-
barrier IPM rotor during high-speed operation. The effects of rotor eccentricity of IPM
machines structures were further investigated by Hwang et al. [89]. El-Refaie et al. [90]
modelled the thermal properties of a multi-barrier IPM machine by the use of the
lumped-parameter and FE methods, showing that these two methods were in good
agreement.
An important factor increasing the reliability of the drive system is its inherent ability to
tolerate faults. Welchko et al. [91] studied the effects of single-phase faults on IPM
machines, and highlighted some useful points in the design of the drive control strategy
to minimise the effects of faults. Bianchi et al. [92] proposed a method to compensate
unbalanced faults in the three phases to maintain smooth torque production.
1.2.2 Concentrated Non-overlapping Windings in AC machines
Up to the early 21
st
century, CW were mostly applied to DC machines. Earlier
published papers on the application of CW in PM AC machines recognised increased
EMF harmonics, lower torque density and narrower CPSR compared to machines with
DW [36]. Later studies by Cros et al. [16, 93], as well as Magnussen and Sadarangani
[16, 17, 93] proved that a winding factor very close to unity and low cogging torque are
achievable if an appropriate slot and pole combination is chosen. These studies sparked
global research interests in fractional slot CWfor use in3-phase PM AC machines.
There were numerous attempts to find the most ideal slot and pole combination to
achieve desired performance characteristics. Nakano et al. [94] illustrated the increases
Chapter 1: Introduction
17
eddy current loss in SPM machines with CW due to asynchronous rotating components
in the MMF waveform. They demonstrated how various slot and pole combinations
resulted in lower losses. The results indicate that eddy current loss decreases inversely
with the number of slots. Reddy et al. [95] carried out a detailed study on losses in SPM
machines with CW and graphed losses with FE analysis for various slot and pole
combinations as well as magnet types. J ussila et al. [96, 97] explained and modelled the
losses in low speed CW machines with several possible slot and pole combinations.
They also proposed a set of guidelines in designing machines with CW. El-Refaie and
J ahns [98] studied the scalability of SPM machines with CW in achieving wide CPSR.
In this study they concluded that a high number of slots and poles is beneficial in
achieving a wide CPSR. Xu and Sun [99] as well as Salminen et al. [100, 101]
presented pull out torque and percentage torque ripple comparisons of various slot and
pole configurations for CW-PM machines. Their results show that torque increases as
the number of slots per pole per phase (S
pp
) increases. Gerada et al. [102] compared
CW-SPM generators with various slot and pole combinations for achieving high
efficiencies and high torque densities.
The key favourable characteristic of CW is its non-overlapping coils, which reduce the
overall amount of copper required and shortens end winding length. Murakami et al.
[103] exploited CWs advantage of non-overlapping coils in the design of high-speed
constant torque IPM machines with reduced size and increased efficiency. Using
different prototypes, they showed that flat, rectangular shaped windings resulted in
shorter end winding length and higher slot-fill factor.
Non-overlapping coils also simplify the stator winding process and permit pre-wound
stator assemblies, thus increasing the slot-fill/packing factor. J ack et al. [104] used
Chapter 1: Introduction
18
compressed copper conductors on separable tooth pieces to achieve a 78% slot-fill
factor. They also showed the significant increase in efficiency and torque density in a
machine constructed by this method. Akita et al. [105, 106] showed performance
characteristics of a machine produced by the joint lapped-core method, which achieved
a 75% fill factor. This was one of the pioneering methods which achieved such a high
slot-fill factor on laminated cores. There were also several patents regarding the
manufacturing of CW stators with the aim of simplifying the winding process and
achieving a high slot-fill factor [107-111].
CW is often contrasted with the more commonly applied DW structure. Asano et al.
[112] showed that CW produced larger radial forces acting on each stator tooth,
resulting in increased vibrations compared to DW. They addressed this issue by
enlarging the airgap size. Magnussen et al. [113] compared parameters, torque and
speed characteristics of an SPM machine having distributed and CW with different
Spp combinations. They concluded that the DW model produced larger torque up to
base speed while the CW model achieved a wider CPSR. Kwon et al. [114] presented
a comparative study on the IPM machine with CW and DW for field-weakening
applications. In this study they showed that the saliency ratio in CW was lower
compared to DW, leading to lower reluctance torque and hence lower overall torque
produced. However, it was also noted that lower current was required for weakening the
field, making the CW machine more efficient in the field-weakening region.
CW are usually classified as single-layer, (alternate teeth wound), or double-layer, (all
teeth wound). Both the single and double-layer windings have advantageous
characteristics in different applications. Ishak et al. [115] compared machine parameters
and performance characteristics of a single and double-layer CW SPM machine. They
Chapter 1: Introduction
19
concluded that the latter resulted in significantly lower torque ripple, but also had
reduced torque density. El-Refaie et al. [18, 116] emphasised the importance of
winding inductance on operating performance and presented an analytical model to
calculate inductance. They compared the self and mutual-inductances of single and
double-layer fractional slot CW, and showed that both inductance values were higher
with single-layer windings due to higher slot leakage and MMF harmonics. El-Refaie
and J ahns also compared the efficiencies of single and double-layer CW in an SPM
machine. They highlighted that double-layer windings result in higher efficiency due to
lower harmonic content compared to single-layer windings, but result in lower overload
torque capability. Wrobel et al. [23] studied the thermal performance of a CW outer
rotor PM machine and indicated the significant increase in temperature of the double-
layer wound machine compared to single-layer one.
The inherent fault tolerance capability of CW makes it an excellent choice for high
reliability applications. Abolhassani and Toliyat [117] exploited the fault tolerant
behaviour of CW and designed a five phase machine, which was proven to have little
performance de-rating under single phase and switch faults. Bianchi et al. [20] presented
design considerations including the fault resistant capability of a fractional slot CW.
They pointed out that while single-layer windings present higher fault resistant
capability, they are more suited for machines with larger airgaps due to the increase in
airgap harmonics. Shah et al. [118] analysed faults in double-layer, vertical and
horizontal concentric wound stators. Fault currents measured from these two
configurations had almost exactly the same magnitudes.
There was also a significant amount of research interest in optimising the performance
of CW machines with various rotor types and structures. Lindh et al. [119] presented
Chapter 1: Introduction
20
detailed comparisons on CW-IPM and CW-SPM motors having open and semi-closed
stator slots. They concluded that the choice of IPM rotors resulted in higher overall
efficiencies compared to SPM rotors with open slots. Salminen et al. [120] also
compared the performance of CW machines between the SPM and IPM rotor types with
CW. Their results indicated that machines with surface magnets produced larger pull
out torque, whereas machines with embedded magnets resulted in lower copper loss.
Han et al. [121] compared losses between a single- and double-layer magnet barrier
rotors with DWand CW. Experiments carried out showed a trade-off between the rotor
and stator losses in the two rotor types. Kawaguchi et al. [122] proposed a cogging
torque reduction technique in an IPM machine with CW by rotor flux barrier shaping.
Kano et al. [123, 124] also attempted to reduce torque ripple in a CW-IPM machine by
varying the rotor flux barrier angle. Kim et al. [125] also produced a design to reduce
torque ripple and cogging torque by simple variations to both the stator and rotor core
geometry. Lee et al. [126, 127] compared the performance of a CWIPM machine with
various rotor shapes, and concluded that C-shaped magnets had the most potential in
reducing torque ripple and improving torque density. Kim et al. [128] studied the
demagnetisation characteristics of various buried rotor types with CW, and stated that
the multi-barrier design was the least susceptible to demagnetisation compared to the
single-layer and v-shaped designs.
The application of CW resulted in an increase in inductance values which helped in
achieving optimal field-weakening conditions. This created significant research interest
in the application of CW to machines used for field-weakening applications. El-Refaie
and J ahns [129] first performed a comparison of different machine types including: IPM
and SPM machines with DW and an SPM machine with CW. The comparison proved
Chapter 1: Introduction
21
that SPM machines with CW had excellent field-weakening potential. Later, El-Refaie
and J ahns [130, 131] exploited the field-weakening potential of CW in SPM machines
and performed an in-depth study in this area of application. They successfully designed
and built a prototype which managed to achieve very good efficiency with an extremely
wide 6:1 experimental CPSR. Soong et al. [132] demonstrated the use of parameter
planes on SPM machines with CW to narrow down the design region, adhering to
freedom car specifications. Munoz et al. [133] presented a comparison of PM machines
with CW and DW. They noted that the benefits of having shorter end turns with CW is
lost as stack length is increased. In terms of field-weakening ability, CW reduces the
saliency ratio thereby limiting the field-weakening ability of the machines. Deak et al.
[134] measured the field-weakening performance of two different machines: one having
double-layer CW with semi-closed slots and the other having single-layer CW with
open slots and unequal tooth-widths. It was shown that overall losses for the single-
layer winding machine were lower in the constant torque region but became larger
compared to the double-layer machine in the wider constant power speed regions.
Although the increase in harmonic content created by CW aids in achieving optimal
field-weakening conditions, it also introduces additional harmonics, thus increasing core
and magnet losses especially at higher frequencies. A significant amount of research has
been conducted to model and reduce these additional losses. Mellor et al. [135]
illustrated a computationally efficient FE method used to calculate losses and power
output from SPM machines with CW. Meier and Soulard [136] attempted to estimate
iron loss from measured magnetic flux densities of an SPM machine with CW. They
illustrated the difficulties in calculating losses in tooth-tips and yoke due to non-uniform
flux distribution. Nuscheler [137] also presented a method to calculate the rotor eddy
Chapter 1: Introduction
22
current losses of PM machines with CW, in which he discussed how losses can be kept
to a minimum. Bottauscio et al. [138] measured losses of a SPM machine with CWat
various operating speeds, considering both solid and laminated rotor cores. They
showed that while PM losses remained relatively constant, there was a huge difference
in rotor core losses (rotor core loss was almost negligible with the laminated core).
Nakano et al. [139] modelled the eddy current losses in a SPM machine with CW and
emphasised the importance of choosing an appropriate slot and pole combination in
achieving low losses. Yamazaki et al. [140] performed a rotor loss comparison between
the IPM, inset and SPM rotors, in which he showed that the IPM had lowest magnet
losses followed by the inset. The SPM produced the greatest amount of losses but could
be effectively reduced by increasing the number of segments. It was also confirmed that
magnet losses in CW are higher compared to DW [141]. Polinder et al. [142], although
not successful in modelling the losses of machines with CW, indicated useful pointers
to consider before loss modelling.
The increased harmonic content produced by CW also resulted in increased vibrations
and acoustic noise during operation. Wang et al. [143] presented a study of resultant
vibrations caused by a modular CW machine. Lee et al. [144, 145] presented a study of
acoustic noise on IPM machines with CW. It was shown that with the optimal slot
opening width, barrier and pole angle, normal forces can be reduced and hence acoustic
noises will lessen. Araki et al. [146] addressed the effects of vibrations in a CW-PM
machine and attempted to reduceits effects by the inclusion of slits on the rotor surface.
Some configurations of CW had asymmetrical structures which resulted in unbalanced
radial forces acting on the rotor. Xu and Li [147] pointed out the effects of radial
unbalanced pull forces caused especially by CW, and charted the unbalanced forces for
Chapter 1: Introduction
23
two different 9-slot motors with CW. It was observed that the choice of slot and pole
combination had a very large effect on unbalanced radial pull force.
Often accurate 3D FE modelling requires large amounts of time and computational
resources. Therefore there have always been attempts at deriving a more
computationally efficient and accurate mathematical representation of the CW model.
J ingai at al. [148] illustrated mathematical models of brushless motors with CW under
various operating conditions. Qu and Lipo [149] presented a general closed-form
analytical method which could be used to model single and double-layer CW.
Abdennadher et al. [150] presented an analytical derivation of inductances of PM
machine with CW and DW. Tangudu et al. [151, 152] presented a lumped parameter
circuit model for an IPM machine with fractional slot CW. They also proposed an
economical and accurate method to separate the reluctance and magnet torque terms.
Meier and Soulard [153] attempted to validify the d-q axis theory on SPM machines
with CW. They showed that steady state operating characteristics such as torque can
still be calculated despite the increase in MMF harmonics. Duan et al. [154] proposed a
particle swarm optimisation method to optimise designs of SPM machines with CW.
This proposed method had a faster computation time compared to FE analysis and was
proven to agree with analytical methods.
Currently there are keen research interests in the sensorless control of PM AC machines
with CW. Reigosa et al. [155] presented a comparative study on the self-sensing
capability of various machine types, including the IPM machine with fractional slot CW.
In this study they made useful non-subjective pointers in which preferred attributes for
both CW and DW were stated. Eilenberger el al. [156, 157] implemented sensorless
control on two different PM machines with single- and double-layer CW. It was shown
Chapter 1: Introduction
24
that the estimated rotor position was almost identical to the actual rotor position with
only slight deviations in both machines. They also tested the overload, short circuit and
field-weakening capability of the single-layer machine, and commended on this
performance for future traction applications. Imai et al. [158] studied the effects of
sensorless control of a CW machine with two different rotor configurations. They
showed that the rotor without flux flowing through the inter-pole link sections achieved
better sensorless performance. Sensorless control strategies have also been successfully
implemented in an IPM machine with CW by Kojima et al. [159].
From 2007 onwards, the use of CW-PM machines began to gain popularity for
industrial applications. Wang et al. [160] studied the CW-IPM machine for undersea
propulsion with the requirement of a high power factor. They indicated with the aid of
the parameter plane the trade-off between CPSR and power factor. Ionel [161]
highlighted the increased use of electrical motors for household appliances in the US,
including the use of CW. Cistelecan and Popescu [162] exploited the advantage CWs
simplified winding structure to design a 10kW synchronous wind generator, in which
they also highlighted the need to use double-layer windings because of the lower MMF
harmonics produced. Kazmin et al. [163] presented a comparative study on various
winding configurations and machine geometries to design an in wheel traction motor.
They found the CW-SPM machine most suitable for this particular application. Alberti
et al. [164] designed, manufactured and successfully tested a 1kW integrated starter
alternator, achieving the desired high starting torque and wide CPSR characteristics.
Germishuizen and Kamper [165, 166] designed a large 150kW railway traction machine,
clearly illustrating the design process and considerations; performance results obtained
from the manufactured machine were also documented. In a recent paper, El-Refaie
Chapter 1: Introduction
25
[167] presented a detailed overview of fractional-slot CW PM machine technology and
its applications, highlighting the large number of major research contributions to this
area. He also stated the challenges faced and proposed future work in this area of
research.
1.3 SCOPE AND ORGANISATION OF THESIS
The general objective of this thesis is firstly to present a feasibility study on the IPM
machine with CW used for field-weakening applications. Subsequently to design,
optimise and construct a CW-IPM machine to achieve a wide field-wakening range with
specific design targets. Finally, with the successful design and experimental verification
of the prototype machine, modifications will be proposed to increase its efficiency, as
well as to scale up the machine. The study, design and experimental verification will
help clarify the advantages and disadvantages of implementing CW in IPM machines,
as well as to demonstrate its prospects in industrial applications requiring high
efficiency over a very wide field-weakening range.
This thesis is organised as follows:
This chapter covers the general backgroundof relevant technology. It also provides an
in-depth review of relevant work for this thesis and highlights the scope and
organisation of this thesis.
Chapter 2 will present numerical techniques in solving electromagnetic field problems.
Of the various numerical methods, the most commonly used method the FE method
will be explained in greater detail. The chapter will show how the mathematical
formulation of the physical model is derived; also how approximations are formulated
Chapter 1: Introduction
26
and solved using Galerkins method and Newton Raphsons iteration. Lastly, chapter 2
will give a basic overview of a 2D FE design process with the use of Magsoft-Flux2D.
Chapter 3 will state advantages and disadvantages of implementing CW on the IPM
machine. It will study various parameter optimisation methods that are used with the
CW-SPM machine; the suitability of these methods will be determined for
implementation on the CW-IPM machine. With the aid of FE models, a suitable slot
and pole combination as well as suitable rotor magnet geometry will be determined.
Chapter 4 will clearly state goals for the design of a CW-IPM machine prototype. The
chosen base layout in chapter 3 will be optimised to achieve a wide field-weakening
range and desired torque/power characteristics. The electromagnetic as well as
structural properties of materials, which affect the performance of the machine will also
be studied. Lastly, chapter 4 will give detailed specifications of the optimal model
which will be constructed for verification purposes. The overall machine parameters and
performance characteristics will be predicted using FE analysis and output
characteristics will be shown.
Chapter 5 will show how CW machines are more susceptible to frequency related losses
compared to DW. Losses in various parts of the final machine model will be determined
with the use of 2D and 3D FE analysis. With the material properties available,
mechanical losses will be calculated. The chapter will also show how losses vary with
material type, as well as with different rotor configurations. Finally the overall predicted
losses at various operating speeds will be used to predict the overall efficiency of the
machine.
Chapter 1: Introduction
27
Chapter 6 will illustrate the control methodology, controller architecture and inversion
technique used to produce the final three-phase inputs to the prototype CW-IPM
machine. It will also compare how the d-axis current trajectory calculated by a widely
used vector control technique for DW-IPM differs from the ideal trajectory obtained
through repetitive testing.
Chapter 7 will conclude and verify the entire design process of the CW-IPM machine in
this work. The manufacturing process of the prototype CW-IPM machine will be stated
and detailed performance characteristics of the prototype will be shown. Lastly the
machines performance characteristics will also be compared to two other equally sized
DW-IPM machines.
Based on the confidence gained from the verification of FE results and lessons learnt
from the design of the first CW-IPM prototype, chapter 8 will propose two different
models, (derived from the prototype with geometrical alterations), to optimise efficiency.
Also the scalability of prototype CW-IPM will be studied and scaled up versions will be
presented.
Chapter 9 will conclude the work presented in this thesis and provide suggestions for
future work to be done on the CW-IPM machine.
28
CHAPTER 2
NUMERICAL METHODS FOR THE ANALYSIS AND PREDICTION
OF MACHINE PARAMETERS
2.1 INTRODUCTION
Present day numerical methods provide a high degree of accuracy in the design of
electrical machines. They allow the designer to predict and optimise machine
parameters and operating characteristics before committing to build the physical
structure, reducing the resources and time required to complete the machine with
desired specifications.
In electrical machines, behaviour is governed by the interaction of electromagnetic
fields created by armature reaction and permanent magnets. These electromagnetic
fields can be expressed in terms of Maxwell equations and the machines performance
can be predicted by solving these equations by various methods, some of which are
shown in Fig. 2.1.
Fig. 2.1 Various methods to solve Maxwell equations and predict machine performance
Numerical
methods
Finite element
method
Electromagnetic
analysis
(Maxwell equations)
Analytical
methods
Lumped-parameter
model
Equivalent circuit
model
Finite difference
method
2D
3D
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
29
The path indicated in bold in Fig. 2.1, shows the preferred method for solving most
modern-day, complex magnetic field problems. Although analytical methods have much
faster solving times compared to numerical methods, they are limited to relatively
simple problems. Compared to other numerical techniques, the FE method has the
following advantages [168]:
Applicable to any field problem (Some of which are: electromagnetic, heat
transfer and stress analysis).
No restriction on geometrical shape.
No restriction on boundary conditions and loading.
No restriction on material properties.
Accuracy versus time can be varied according to the choice of mesh size.
The FE solution closely resembles the actual field distributionin the region.
Chapter 2 will give a general overview of the FE method, leading up the FE design
methods used into this work. It shows how mathematical formulations of the physical
model are derived, as well as how approximations are formulated and solved using the
Galerkins method and Newton Raphson iteration.
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
30
2.2 THE FINITE ELEMENT METHOD
The FE method is a numerical technique used to solve field problems of structures with
complex geometries and non-linear material properties. This is done by dividing the
entire region of analysis into smaller finite elements which are connected to the
neighbouring elements by two or more nodes. By applying field equations in each of the
elements and ensuring continuity at each node, equations for the entire structure can be
solved. FE analysis of a physical event can be typically expressed in the following steps:
Pre-processing
Construction of geometry
Subdividing study area into finite elements (Meshing)
Assigning material properties
Assigning excitation sources
Assigning boundary conditions
Solving
Deriving and assembling element matrix equations
Solving the equations for unknown variables
Post-processing Analysis of results obtained
There are numerous forms of finite element applications, some of which are: structural,
magnetic and thermal. These applications can be classified into different types of field
problems, such as:
Static
Input sources are time invariant
Used for steady state DC/fixed load analysis
Transient
Input sources are time-variant
Used for transient AC/load-variant analysis
Steady State
Field sources are time-variant but periodic
Used for steady state AC/constant load-variation analysis
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
31
2.2.1 Brief Background of Finite Element Analysis
The first use of FE analysis for solving magnetic field problems involving rotating
electrical machinery dates back to the 1960s. Ahamed and Erdelyi [169] introduced a
mathematical model for the salient pole machine in terms of partial differential
equations. Erdelyi et al. [170] then implemented the theory based on this method to
create an FE model to model the transient behaviour of a DC motor. Silvester and Chari
[171, 172] were among the first to introduce solutions to non-linear electromagnetic
field problems for electrical machines. Since then, there has been ongoing research in
FE analysis of electrical machines. Although the FE method, (like all other modelling
methods), cannot be a hundred precent accurate, its ability to solve complex problems
with high degrees of accuracy has made it an essential part in the design of almost all
high performance electrical machines.
2.2.2 Mathematical Formulations of the Physical Model
Maxwells equations are used to describe the non-linear magnetic fields within the
electrical machine and how they relate to their sources. In differential form Maxwell
equations can be expressed as:
v

= 0
(2.1)
v

= [
(2.2)
v

= -
oB

ot
(2.3)
The zero divergence of magnetic flux (B

) in (2.1) indicates that no source is present.


(2.2) describes how magnetic field (E

) reacts to the electric current density ([). (2.3)


represents the induction of electric field (E

) due to the negative rate of change of


magnetic flux.
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
32
Mathematical formulations for the physical model involve solving the electromagnetic
problem. This is done by solving for the magnetic field distribution given an input
source that is, solving (2.2). For simplicity, only the current source excitation is
considered first. In this case, the magnetic field is proportional to the flux density as
expressed in (2.4):
B

= pE
(2.4)
where,
p =Permeability of the material
By substituting (2.4) into (2.2), we get:
The electromagnetic problem can simplified by introducing the magnetic vector
potential (A

) and electric scalar potential (u). The relationship between these two
potential values and field variables is shown in (2.6) and (2.7) respectively.
B

= v

(2.6)
E

+
oA

ot
= -v

m
(2.7)
Substituting (2.6) into (2.5):
1
p
v

(v

) = [ (2.8)
Simplifying this problem further with the Laplacian of A

, (2.8) can be expressed as:


1
p
|v

(v

) -v

2
A

] = [ (2.9)
and subsequently simplified by the use of Coulomb gauge (v

= 0) to give:
-
1
p
v

2
A

= [ (2.10)
v

1
p
B

= [ (2.5)
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
33
Current density can be expressed in terms of field values shown in (2.11) and the field
values can then be replaced by potentials from (2.6) and (2.7).
[ = o(E

+v B

) (2.11)
[ = o _-v

m-+
oA

ot
+v (v

)_ (2.12)
where,
o =Conductivity of the material
v =Velocity of a material with respect to a given reference frame
From (2.10) and (2.12):
-
1
p
v

2
A

+o
oA

ot
-v o(v

) = -ov

m
(2.13)
(2.13) is the governing equation physical model without permanent magnet excitation.
To create a model involving permanent magnet excitation, the permanent magnet
material can be modelled by a magnetisation vector model based on the B-H curve.
Considering permanent magnet excitation, (2.4) becomes:
B

= p
0
(E

+H

) (2.14)
where,
H

=Magnetisation vector
By substituting (2.14) into (2.2) the equation can be modified to include permanent
magnet excitation in addition to the current only excitation:
-
1
p
v

2
A

+o
oA

ot
-v o(v

) = -ov

m+v

H
(2.15)
The term ov

m describes the input resulting from armature current and v

describes
input resulting from permanent magnets. For the sake of simplification these two terms
will be written as J
a
and J
m
respectively. To eliminate the velocity term, the moving
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
34
frame is taken as the reference frame and the relative velocity becomes zero. (2.15) is
then reduced to:
-
1
p
v

2
A

+o
oA

ot
= -J
a
+J
m
(2.16)
(2.16) is the general governing equation for the physical model with time varying fields
with the consideration of both current and permanent magnet excitation.
In two dimensions, the magnetic vector potential has only z-axis components, thus it
can be reduced to the z-axis potential ( A
z
). The governing equation for a two
dimensional analysis is therefore given by:
-
1
p
v

2
A
z
+o
oA
z
ot
= -J
a
+J
m
(2.17)
The partial differential equation in (2.17) can be regarded as the system equation. With
the magnetic vector potential A
z
used as a key variable, the entire study domain can be
discretised by a mesh of finite elements.
2.2.3 Discretisation of the Study Domain
The FE method solves the partial differential equation (2.17) by looking for an
approximate function which describes the behaviour of each element. The elements are
held together by nodes. The accuracy of the solution would depend on the size of the
elements. A smaller mesh results in higher accuracy, but requires more elements to
make up the study domain, leading to a longer computing time. In general, the mesh
should be most dense where field-change gradients are the largest and least dense where
the field change is relatively constant. Fig. 2.2 shows some typical elements for one-,
two- and three-dimensional problems. For one-dimensional domains, shorter
interconnected line segments make up the original linear structure. For two-dimensional
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
35
domains, common elements would be rectangles for discretising rectangular regions or
triangles for more irregular-shaped regions. For three-dimensional domains,
tetrahedrons and hexahedrons are the more commonly used element shapes.
One dimensional
(line) elements
Two dimensional
(surface) elements
Three dimensional
(brick) elements
Fig. 2.2 Typical finite elements
Each of these elements can be represented by a polynomial in terms of the potential A.
For example the basic three node triangle with nodal degrees of freedom o
n
can be
represented by:
A = o
1
+o
2
x+o
3
y
(2.18)
For a six node triangle, the expression would be represented by a quadratic equation
(2.19) with six nodal degrees of freedom.
A = o
1
+o
2
x +o
3
y +o
4
x
2
+o
5
xy +o
6
y
2
(2.19)
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
36
For a three-dimensional brick element, the field can be expressed as:
A = o
1
+o
2
x +o
3
y +o
4
z +o
5
xy +o
6
yz +o
7
zx +o
8
xyz (2.20)
Solving these polynomials requires the constants o
n
to be solved in terms of the
interpolated field variables A
n
.
Using the basic three node triangle as an example, unknown constants o
1
, o
2
and o
3
have to be solved. The three node triangle is represented by x and y-axis coordinates
with x
1
= y
1
= y
2
= 0 as shown in Fig. 2.3.
Fig. 2.3 Two dimensional triangular element
Evaluating the expression at the nodes:
_
A
1
A
2
A
3
_ = _
1 0 0
1 x
2
0
1 x
3
y
3
_ _
o
1
o
2
o
3
_
(2.21)
Solving for o
n
:
o
1
= A
1
o
2
=
A
2
-A
1
x
2
o
3
=
(x
3
-x
2
)A
1
x
2
y
3
-
x
3
A
2
x
2
y
3
+
A
3
y
3

(2.22)
Thus the interpolated field expression for the three node triangle is given by:
A = [C][P]
-1
{o} (2.23)
A
1
(x
1
, y
1
)
A
2
(x
2
, y
2
)
A
3
(x
3
, y
3
)
)
x-axis
y-axis
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
37
where,
[C] = [1 x y][P]
-1
[P]
-1
=
l
l
l
l
l
1 0 0
-1
x
2
1
x
2
0
x
3
-x
2
x
2
y
3
-x
3
x
2
y
3
1
y
3
1
1
1
1
1
{o} = _
A
1
A
2
A
3
_ (2.24)
The discretised magnetic vector potential in each element can be generally expressed as:
where,
J =Number of nodal degrees of freedom
Unknown field quantities in (2.17) can then be replaced by (2.25). These replaced
quantities however, contain errors. These errors or residuals can be reduced if error
minimisation techniques are used to derive the equations. Of the many error
minimisation methods available, Galerkins method is most commonly used for solving
boundary value problems in electromagnetics [173]. Before applying Galerkins method,
certain boundary conditions have to be stated. These boundary conditions will be shown
in the next section.
2.2.4 Definition of Boundary Conditions
Imposed boundary conditions affect the accuracy and efficiency of the FE solution
[174]. There are three main groups of boundary conditions:
Dirichlets condition
Neumanns condition
Periodicity condition
A

= A

n
C
n
(x, y)
d
n=1
(2.25)
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
38
Dirichlets Boundary Condition
In Dirichlets boundary condition, a value A

is assigned on the boundary (I). These


conditions are most commonly imposed on the surface of the rotor inner diameter C-D
and stator outer diameter A-B as shown in Fig 2.4.
Fig. 2.4 Dirichelet boundaries
It is common to assign the condition of A

= 0, as long as the leakage flux beyond these


boundaries is negligible. The high magnetic permeability of the core materials normally
ensures that the majority of the flux is contained within the boundaries, hence making
the assumptions valid.
Neumanns Boundary Condition
Neumanns boundary condition is satisfied when the derivative of A

is assigned to be
normal to theboundary (I). Neumanns boundary condition is normally imposed to a
region that has symmetry.
A
B
C
D
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
39
Periodic Boundary Condition
Periodic boundary condition is assigned between two or more boundary lines using
symmetries in the geometry or magnetic field distribution. Among these lines, a
principal line is selected and the other line(s) is (are) expressed as a function of the
potential of the principal line. This boundary can significantly reduce the size of the
numerical model. Fig. 2.5 shows a 4-pole machine with quarter symmetry..
Fig. 2.5 Machine with quarter cyclic-symmetry
It is advisable to solve boundary value problems analytically whenever possible.
However, most practical problems do not have an analytical solution, so popular
approximate methods such as Galerkins method are used to derive the FE equations.
2.2.5 Galerkins Method for Deriving Finite Element Equations
Weighted residual and variational methods are the more commonly used methods in
deriving FE equations. Galerkins method is a particular form of the weighted residual
method and is the most popular method due to its accuracy and simplicity of
C
A
B
D
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
40
implementation. Galerkins method deals with the differential equation directly and
solves the field problem by residual reduction [175-177].
To describe Galerkins method, a domain is declared, in which the physical problem
is defined, letting I(A
-
(x, y)) be a differential operator containing the trial solution
A
-
(x, y) and (x, y) be a given forcing function of x and y.
The trial solution A
-
(x, y) will contain an error and there remains a residual (R(x)).
The mathematical statement to the physical problem can be expressed as:
In domain : I(A
-
(x, y)) -(x, y) = R(x, y) (2.26)
By the use of a weight function (w

), the integral of the residuals is forced to zero over


the domain. This can generally be expressed as:
w

R(x, y)Jx Jy = 0

(2.27)
In terms of the machine model equation in (2.17), the expression would be:
_-w

1
p
v
2
A
z
+w

o
oA
z
ot
+w

J
a
-w

J
m
] Jx Jy = 0

(2.28)
For Galerkins method, w

in each element is chosen to be equal to an interpolating


function (I

) [178]:
w

= I

for i =1, 2 , 3, , E (2.29)


where E is the total number of elements.
The general expression in computational matrix form is given by:
[S]{o} = {[}
(2.30)
where[S] is the global coefficient matrix which is dependent on interpolating functions
and its elements in the i
th
column and ]
th
row S
]
are given by:
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
41
S
]
= (I

II

1
)

JxJy
(2.31)
As before {o} is the columnvector of unknown coefficients:
{o} = _
A
1
A
2
A
3
_ (2.32)
and{[} is the column vector whose elements are dependent on the forcing functions (J).
[

= (I

J)

JxJy
(2.33)
In electromagnetics, (2.30) describes a deterministic system where there exists an input
source or excitation.
2.2.6 Solving Finite Element Equations with the Newton Raphson Method
Solving electric machine problems involves dealing with non-linearities in both
electromagnetic and structural material characteristics. For this thesis, an example of a
problem involving non-linearity is calculating mechanical stress inflicted on the inter-
pole link sections under high speed operations, where the stress on the core material is
not directly proportional to the strain. These non-linear equations can be solved using
iterative methods such as the Newton Raphson method [179, 180] a common method
used for solving FE problems.
To formulate the Newton iteration, a general 2D energy functional (F) will first be
defined. From (2.9), where the magnetic potential is expressed in terms of the input
current density, the following expression can be obtained:
o
ox
_
1
p
oA
ox
] +
o
oy
_
1
p
oA
oy
] = -[ (2.34)
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
42
and a suitable functional for (2.34) would be:
F(A
-
) = w(A
-
)JxJy -[

A
-
JxJy
(2.35)
where w(A
-
) is the energy density associated with the trial solution.
With the determined functional F(A
-
) and the trial solution A
-
, the gradient of F(A
-
) is
first found by Taylor series expansion:
oF
oA
-

=
oF
oA
-

_
A
-
+
o
2
F
oA
-

oA
-
]
_
A
-
d
]=1
_
o
2
F
oA
-

oA
-
]
_
A
-
_
-1
_
oF
oA
-
]
_
A
-
_ + (2.36)
The iteration process can now be set up, letting A be the correct solution to A
-
. The trial
solution can be expressed as:
A
-
= A +_
o
2
F
oA
-

oA
-
]
_
A
-
_
-1
d
]=1
_
oF
oA
-

_
A
-
_ (2.37)
With an initial assumption of the estimated potential A
-
, the iterative process can begin
and the step difference from the correct potential A can be calculated. This step
difference is added to A
-
to achieve a better estimate and the entire process is repeated
until the correct potential is found. Fig. 2.6 gives a basic procedure for Newton
Raphsons method of achieving convergence:
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
43
Fig. 2.6 Newton Raphson method
This iterative process can be generally expressed as:
A
-
(n+1)
= A
-
(n)
-_
o
2
F
oA
-

oA
-
]
_
A
-
_
-1
(n)
d
]=1
_
oF
oA
-

_
A
-
_
(n)
(2.38)
2.2.7 Process of a Time Stepping Finite Element Model
When solving a time-dependent problem, the solution to the whole problem consists of
solving the problem at discrete time intervals (time steps). The solutions at each time
interval are dependent on the solution of the previous time step. Of the various
application types, transient or time stepping analysis is most commonly used for the
analysis of electrical machines, due to many time-dependent aspects such as: movement
of the rotor, time-variant excitation of individual phases and the developed torque. The
flowchart for the time stepping solving process is shown in Fig. 2.7:
F(A
-
)
A
-
00
A
-
1
A
-
2
A
-
3
A
-
(n)
F(A
-
)
Slope
P
A
-
0
Slope
P
A
-
1
F(A
-
0
)
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
44
Fig. 2.7 Flow chart for the time-stepping finite element process
Yes
Terminate time
stepping
End
No
Yes
Yes
Time Step
(T
step
=T
step
+1)
Rotation of
rotor mesh
Generation of
coefficient matrix
Start
Error
(solution)
<
Error
(tolerance)
Post Process
Recalculate in
next iteration
Solve for
approximate solution
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
45
2.3 FINITE ELEMENT METHOD FOR DETERMINING MACHINE
PARAMETERS AND PERFORMANCE CHARACTERISTICS
In this work, a number of machine models were created, compared and optimised FE
analysis. The final design chosen for prototype construction is a concentrated wound,
double-layer, 18-slot, 14-pole, IPM machine with v-shaped magnets, as shown in Fig.
2.8. Here, the design process of the model using Magsoft-Flux2D will be discussed in
detail.
Fig. 2.8. Prototype machine geometry showing regions
The stages in a basic design process are as follows:
Construction of geometryand assignment of mesh (discretisation)
Definition of material properties
Performance analysis
2.3.1 Construction of Geometry and Assignment of Mesh
During the construction of geometry, it is important for the designer to recognise
symmetry in the design. When used together with the appropriate boundary condition,
the computational time can be significantly reduced due to a reduced size of the study
Stator
core
Stator
windings
Rotor
magnets
Rotor
core
Shaft
Airgap
Stator slot
opening
Air sections
in magnet
slots
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
46
domain. Common types of symmetry include axis symmetry and plane symmetry. For
rotational machines, axis symmetry can be applied if the revolutionary axis passes
through the origin of the coordinate system. For the models created in this work, axis
symmetry (quarter symmetry in particular) could only be used on the integral-slot DW
machines, where the phase coils are equally distributed around the stator (fig. 2.9 shows
a model of UNSW segmented DW-IPM with phase coils repeating every quarter).
Fig. 2.9. DW-IPM showing the repletion of phase coils every quarter
After geometrical construction of the principal portion, a mesh has to be created. In FE
analysis, the accuracy of the approximate solution depends to a great extent on the
quality of the mesh. Some general rules for meshing include:
The mesh should be well proportioned.
Ideal elements for a surface mesh should equilateral triangles or squares.
The mesh should not be unnecessarily fine.
A typical mesh structure for the prototype machine with consideration of the
abovementioned rules is shown in Fig. 2.10.
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
47
Fig. 2.10 Mesh structure of the CW-IPM machine
After the mesh is created, the quality of mesh can be further validated by observations
such as:
The field lines should be continuous within a specified region.
The action force should be equal to the reaction force in a rotatingmachine.
Current and voltage through a coil should be the same when computed by
different methods.
2.3.2 Defining Material Properties
Before defining the desired regions to parts of the created machine geometry, material
properties have to be defined for example, properties for the permanent magnets, rotor
and stator core, as well as windings. Material properties are made up of various
behavioural laws such as:
Material Type Constitutive Equations Material property
Magnetic
B

= pE
(2.4) p =Magnetic permeability
Dielectric

= eE
(2.39) e =Electric permeability
Conductive
[

= oE
(2.40) o =Electric conductivity
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
48
Additionally, materials may be anisotropic and non-linear; material properties may also
vary with temperature and frequency. Thus, the behavioural laws can get very complex
if an accurate representation of materials is required.
Magnetic materials are generally characterised by the hysteresis cycle, otherwise known
as the B-H curve. B-H curves for soft and hard magnetic materials are shown in Fig.
2.11a and 2.11b respectively.
Soft Magnetic Materials
(Core)
Hard Magnetic Materials
(Permanent magnets)
(a) Soft magnetic materials (b) Hardmagnetic materials
Fig. 2.11 B-H Curve for hard and soft magnetic materials
Isotropic electrical steel laminations are chosen as the core material for this prototype
and an analytic solution is used to model its properties. The solution consists of a
straight line representing the linear region (2.39) and an arctangent curve (2.40)
representing the saturated region shown in Fig. 2.12. This solution is used instead on
tabulated B-H curve as specific material properties were not availableat the beginning
of this work. Furthermore, having an exact material match was not crucial to the
outcome of this thesis.
B
H
1
st
Quadrant
3
rd
Quadrant
B
H
2
nd
Quadrant
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
49
Fig. 2.12Representation of core material characteristics in terms of an analytic solution
B(E) = p
0
p

E (2.39)
B(E) = p
0
E +
2[
s
n
orcton _
n(p

-1)p
0
E
2J
s
_ (2.40)
where,
p
0
=magnetic permeability of free space
p

=initial relative permeability


J
s
=magnetic polarisation at saturation
Permanent magnets are chosen as the rotor field source. The performance of magnets is
most commonly described in the second quadrant of the B-H curve [181]. The magnet
temperature rating is chosen such that the knee of the demagnetisation curve doesnt
enter the second quadrant, and the B-H curve can be represented by a linear function, as
shown in Fig. 2.13. This unidirectional linear function can be represented by (2.41).
Fig. 2.13Representation of permanent magnet material demagnetisation characteristics
B [T]
H [A/m]
Linear
representation of
unsaturated region
p
0
p

Asymptote line passing through J


s
B = p
0
E +J
s
J
s
Saturated region represented
by an arctangent curve
B [T]
H [A/m]
B

t
E
c
Demagnetisation
line p
0
p

Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
50
B
//
(E) = p
0
p
//
E
//
+B

(2.41)
After declaration of B-H characteristics, the magnet orientation has to be defined.
Common choices of orientation include positive radial, negative radial, positive
orthoradial and negative orthoradial. The prototype, however, has v-shaped magnets,
thus making the abovementioned orientation choices invalid. The magnets shown in Fig.
2.8 have to be individually orientated.
2.3.3 Coupling of Electrical Circuits
An external electric circuit can be coupled to the magnetic model via solid or stranded
conductors. While solid conductors reduce I
2
R losses at low frequencies, this form of
conductors result in high eddy current loss at higher frequencies, thus stranded
conductors are more suitable for AC machine designs. A typical circuit with three-phase
current excitation is shown below, (noting that a two phase current excitation is used to
represent the actual three phase model in the mesh current method [182]):
Fig. 2.14 Three-phase star conneted circuit with current excitation
End winding
inductances
Current sources
Positive and negative
phase coils
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
51
Current sources are defined, and coils of the three different phases are assigned to their
respective regions in the FE model. The terminal voltage (I) comprises the voltage
generated due to the winding resistance (R), end turn inductance (I) and generated back
EMF due to flux () as show below:
I = iR +I
Ji
Jt
+
J
Jt
(2.42)
2.3.4 Performance Calculation
Back EMF
To determine the induced voltage in the machine, a prime mover is required. In
Magsoft-Flux2D this can simply be done by assigning a fixed rotor speed. A similar
circuit to Fig. 2.14 can be used, except that in this case the current sources would be
replaced by infinite resistance assigned to all three phases. Without excitation, the
induced voltage in the conductors (2.43) would only be from the instantaneous induced
back EMF (c) given by:

c =
J
Jt

(2.43)
Cogging Torque
The interaction between the rotor magnets and the stator teeth results in a variation of
reluctance in the airgap. This amount of variation, as well as the magnitude of magnet
flux linked across the airgap, determines the magnitude of cogging torque produced. In
Magsoft-Flux2D, cogging torque (I
cog
) is found by rotating the rotor slowly with an
imposed speed and finding the rotational force (F
ot
) using the virtual works method, as
expressed in (2.44).
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
52
F
ot
= -
ow
os
ds

(2.44)
where,
w =Mechanical virtual work done
(s
ds
) =Mechanical displacement
For the calculation of cogging torque, the airgap mesh should be made up of very good
quality elements. Thus it is advisable to use a three layer airgap as shown in Fig. 2.15:
Fig. 2.15. Meshing of a three-layer airgap
Output Torque
In Magsoft-Flux2D, generated torque can also be calculated by the virtual work method
as described earlier. The circuit in Fig. 2.14 can be used and excitation current values
can be defined as:
i
u
= I
ms
V2(sin 2n t +y)
(2.45)
i
b
= I
ms
V2 _sin 2n t +
2n
3
+y] (2.46)
where,
=frequency of excitation
t =time at each iteration
y =the current angle
Compressible
airgap
Stationary
airgap Rotating
airgap
Chapter 2: Numerical Methods for the Analysis and Prediction of Machine Parameters
53
2.4 CONCLUSION
This chapter introduced the FE method used for modelling machine characteristics and
performance. It presented the different stages and types of field problems of FE analysis
relevant to this work. The derivation of the system equations in terms of partial
differential equations and subsequent discretisation of the study domain were shown.
This chapter showed various element types and how elements were connected through
expressions evaluated at nodes. Galerkins method for deriving the FE equations and
Newton Raphsons method of achieving an error-free solution were shown. Lastly, this
chapter also stated the FE design process of the prototype machine, indicating how key
machine, parameters and performance characteristics were calculated in Flux2D.
FE methods shown in this chapter will be used in this thesis to compare, design and
optimise permanent magnet machine models. Other than Flux2D
,
other FE methods
used include: ANSOFT-Maxwel13 used to model 3D FE electromagnetic analysis and
ANSYS used to model mechanical stresses in the rotor.
54
CHAPTER 3
INVESTIGATION OF THE CONCENTRATED WINDING IPM
MACHINE FOR WIDE FIELD WEAKENING APPLICATIONS
3.1 INTRODUCTION
The principal cause of developed torque in PM machines is the interaction of the fields
from the permanent magnet and rotating stator electromagnetic field. Therefore the
quality of torque produced is largely dependent on the winding design and the
configuration of the magnet poles.
In three-phase synchronous AC machines, windings which result in sinusoidal back
EMF waveforms are desirable for achieving high efficiency and low torque ripple [183].
Before the 21
st
century, the means of achieving sinusoidal back EMF was primarily to
design windings so that MMF waveforms were close to sinusoidal. This made
distributed windings (DW) the winding of choice until recently when Cros and
Viarouge [16]; Magnussen and Sadarangani [17] proved that sinusoidal EMF in SPM
machines could be generated by the use of fractional-slot CW.
This chapter studies the suitability of applying fractional-slot CW in IPM machines to
achieve the desired performance characteristics. With the help of FE analysis, several
CW models will be created and compared to an integral-slot DW model. Subsequently,
the field-weakening performance of two CW-IPM with different magnet geometry will
be compared to the CW-SPM machine. This chapter is a preliminary study to ensure
that the benefits of applying CW outweigh the disadvantages, so that further research
does not lead to futile designs. It will also set the basis for the design and optimisation
of the final model in the next chapter.
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
55
3.2 CHOICE OF SLOT AND POLE COMBINATION
Despite numerous advantages over DW, such as a higher slot-fill factor, shorter end-
windings and decreased manufacturing complexity, CW has been unpopular for use in
AC machines due to its characteristic of producing MMF and EMF waveforms which
are rich in harmonics. However, in recent years it has been verified that CW has the
capability of producing sinusoidal EMF waveforms with appropriate slot and pole
combinations [16, 17], sparking global research interest in this winding type for AC
synchronous machines.
Slot and pole combinations are commonly referred to by the number of slots per pole
per phase (S
pp
). For example, a 3-phase machine with 6-slots and 2-poles would have an
integral-slot distribution of S
pp
=1, whereas 6-slots, 4-poles would have a fractional
distribution of S
pp
=1/2. This fractional distribution is often termed fractional-slot
windings. Unlike in integral slot windings, the phase coils are not periodic over a pole
pitch, (as seen in fig. 3.1a). Thus, an appropriate selection of phase coil arrangement is
essential in producing a sinusoidal EMF waveform of the desired magnitude. If the S
pp
is too small (as in the case where S
pp
=1/7, shown in fig. 3.1b), a large portion of a rotor
pole of opposite polarity will be under the same stator pole, thus supressing the total
EMF generated in that phase. Fig. 3.1 compares the integral-slot DW machine with
fractional-slot CW machines having 1/7, 2/7 and 3/7 S
pp
.
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
56
(a) 2S
pp
DW (b) 1/7S
pp
CW
(c) 2/7S
pp
(d) 3/7S
pp
CW
Fig. 3.1 14-pole IPM machine with various slot and pole combinations
While selecting the appropriate the S
pp
to get the desired EMF waveform, the periodicity
between the stator teeth and rotor magnets also changes. Consequently, this changes the
peak cogging torque value. The following sub-sections illustrate the effects of these two
parameters (EMF and cogging torque) and how they can be estimated by simple
analytical methods.
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
57
3.2.1 Winding Factor and EMF
In an IPM motor, torque is generated by the interaction between the permanent magnet
field and the stator electromagnetic field caused by the armature reaction. The
developed torque for a single phase excitation can generally be expressed as follows:
I
out
(i
s
, 0

,
pm
) =
1
2
i
s
2
JI
s
J0

-
1
2

pm
2
JJ
J0

+N
ph
i
s
J
pm
J0


(3.1)
where,

pm
=Flux from permanent magnets
i
s
=Supply current (to the stator coils)
The first term in the torque equation, extracted to (3.2), is known as the reluctance
torque (I
cI
), and is related to the variation of stator inductance (I
s
) with rotor position
(0

). This variation of stator inductances is caused by pole saliencies or flux barriers.


Reluctance torque is contributes to the total torque generated by the IPM machine.
I
cI
(i
s
, 0

) =
1
2
i
s
2
JI
s
J0


(3.2)
The second term, shown in (3.3) is known as cogging torque (I
cog
). It caused by the
interaction of magnet flux and stator slots, resulting in the variation of reluctance (J).
I
cog
(
pm
, 0

) = -
1
2

pm
2
JJ
J0


(3.3)
The last term, shown in (3.4), is known as electro-dynamic torque or alignment torque
(I
c
). This is the main contributor to the useful torque produced by the machine. I
c
is
related to the variation of mutual inductances between the stator and rotor. In particular,
it is the interaction between the back EMF and stator current.
I
c
(i
s
,
r
,
pm
) = k
w
N
ph
i
s
d
pm
d
r
(3.4)
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
58
The winding factor (k
w
) affects the shape and magnitude of flux linkage across the air
gap, and affects how harmonics of individual coil back EMF phasor are summed
together to form the overall phase EMF. A low winding factor means that the harmonic
components of the EMF are relatively high as compared to the fundamental component
resulting in lower magnitude of useful EMF. A lower winding factor can be treated as a
reduction inthe effective number of turns per phase in stator windings as shown in the
electro-dynamic torque equation (3.4). The EMF in turn affects the efficiency and
torque density of the machine. It is therefore important to obtain a high winding factor
in the initial design stages with an appropriate choice of S
pp
.
Generally, the winding factor is made up of three parts:
k
w
= k
p
k
d
k
s

(3.5)
where,
k
p
=Pitch factor,
k
d
=Distribution factor
k
s
=Skew factor
Concentrated windings are usually not skewed, as it will significantly increase the
complexity and cost of constructing the machine. Therefore the third term can be
omitted. Although the pitch factor can be easily calculated, calculating the distribution
factor can be quite complex. An alternative method that can be used to determine both
the pitch and distribution factor together is the EMF phasor method [17]. A simple
example of this method is shown below. Each EMF phasor is represented by:
E

x
= c
][
n2p
s
x

(3.6)
where,
E

x
=Reference EMF phasor element,
x =winding element number,
p =number of pole pairs
s =number of slots.
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
59
A 6-slot, 4-pole machine layout (shown in fig. 3.2) is used in this example.
(a) 6-slot, 4-pole design showing phase
coils and winding element numbers
(b) Phasor diagram showing how each element
phasor make up for each phase
Fig. 3.2 EMF phasor diagram
By substituting element numbers (x =0,1,2,..5) into (3.6), and by Eulers equations we
get:
For x =0:
E

0
= c
][
n4
6
0

E

0
= cos(0) +]sin(0) = 1zu
For x =1:
E

1
= c
][
n4
6
1

E

1
= cos(-0.5) +]sin(0) = 1z12u
and so on (till x =no. of slots 1).
Vectors E
A/B/C
i
are achieved when all the winding elements corresponding to a
particular phasor are in phase. The winding factor of each phase can then be calculated
by dividing vectorsE
A/B/C
by E
A/B/C
i
respectively.
Although a unity winding factor is not achievable for fractional-slot windings, a value
very close to unity can be achieved with certain combinations. Table 3.1 shows the
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
60
winding factors of single- and double-layer windings with several slot and pole
combinations.
Table 3.1
Winding factor and LCM for different S
pp
values
Poles
4 6 8 10 12 14 16
Slots
6
S
pp
1/2 1/4 1/5 1/7 1/8
k
w (Single layer)
0.866 0.866 0.866 0.866 0.866
k
w (Double layer)
0.866 0.866 0.500 - -
LCM 12 24 30 42 48
9
S
pp
1/2 3/8 3/10 3/12 3/14 3/16
k
w (Single layer)
0.866 0.945 0.945 0.866 0.945 0.945
k
w (Double layer)
0.866 0.945 0.946 0.764 0.473 0.175
LCM 18 72 90 36 126 144
12
S
pp
1/2 2/5 2/7 1/4
k
w (Single layer)
0.866 0.966 0.966 0.966
k
w (Double layer)
0.866 0.933 0.933 0.866
LCM 24 60 84 48
15
S
pp
1/2 5/12 5/14 5/16
k
w (Single layer)
0.866 0.808 0.866 0.866
k
w (Double layer)
0.866 0.906 0.951 0.951
LCM 30 60 210 240
18
S
pp
1/2 3/7 3/8
k
w (Single layer)
0.866 0.945 0.945
k
w (Double layer)
0.866 0.902 0.931
LCM 36 126 144
21
S
pp
1/2 7/16
k
w (Single layer)
0.866 0.932
k
w (Double layer)
0.866 0.851
LCM 42 336
24
S
pp
1/2
k
w (Single layer)
0.866
k
w (Double layer)
0.866
LCM 48
3-phase CW machines with double-layer windings require slots in multiples of 6 in
order for the phase coils to be equally distributed. Thus 6, 12 and 18 slots are chosen
and the same 14-pole rotor will be used.
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
61
Here the back EMF fromthe four designs shown earlier in fig. 3.1 is compared.
(a) 2S
pp
DW
(b) 1/7S
pp
CW
(c) 2/7S
pp
CW
(d) 3/7S
pp
CW
Fig. 3.3 3-phase EMF waveforms and corresponding frequency spectrum over 1 electrical cycle
for 14-pole IPM machine with different slot and pole combinations
0
10
20
30
0 5 10 15
Volt
0
10
20
30
0 5 10 15
Volt
0
10
20
30
0 5 10 15
Volt
0
10
20
30
0 5 10 15
Volt
28.4V
21V
27.6V
14.5V
( )
Time (s) Frequency components (n-order)
Time (s) Frequency components (n-order)
Time (s)
0 5 10 15
Frequency components (n-order)
Time (s)
0 5 10 15
Frequency components (n-order)
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
62
By comparing the back EMF waveforms from fig. 3.3, it can be seen that all, except the
6-slot model, produced near-perfectly sinusoidal (line to neutral) back EMF waveforms.
The 84-slot, DW-IPM machine, (with a calculated winding factor of 0.933), generated
an EMF waveform with the highest magnitude of 28.4V
rms
. From the values shown in
table 3.1, the 12-slot, 14-pole model also has a winding factor of 0.933. Therefore it
should produce the same back EMF magnitude. However, due to the rotor magnet pole
pitch being smaller than the stator pole pitch, the flux from the opposing magnet pole
cancels the total amount of flux that is linked to a phase coil, resulting in a lower than
expected winding factor. The same condition was seen in the 6-slot, 14-pole model.
This indicates that having a larger number of slots than poles are required.
Based on back EMF obtained by the DW-IPM machine, the 18-slot design, which has a
winding factor of 0.902 (from table3.1), the calculated back EMF should by 27.5V
rms
.
From the 18-slot FE model, it is shown that the back EMF magnitude achieved
(27.6V
rms
) agrees with the calculated value.
3.2.2 Cogging Torque
Cogging torque is a result of the variation of stator inductances due to the interaction of
rotor and stator poles. It contributes to vibrations and noise, leading to severe
restrictions in the machine performance. The cogging torque equation was stated in
earlier in (3.3).
There are several methods by which cogging torque can be reduced, some of which are
rotor flux barrier shaping [122], stator chamfer angle shaping [125] and shaping of the
rotor outer radius [61]. These methods, are however, heavily based on trial and error
techniques to achieve the optimal geometry. Another common method to reduce
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
63
cogging torque in machines with integral slot windings is by skewing. However, in
fractional-slot windings skewing would lead to a significant decreasein induced EMF,
as well as increase the complexity and cost of constructing the machine. In actuality, the
use of fractional-slot windings itself aids in the reduction of cogging torque as it
eliminates periodicity between slots and poles.
To quantify the amount of reduction with different slot and pole combinations, the
lowest common multiple (LCM) method is used, which is basically the LCM of the
number of slots and poles. A higher LCM would yield a lower peak value in the
cogging torque waveform but result in higher frequency fluctuation. The LCM of
various combinations is stated in table 3.1. Fig. 3.4. gives an example of cogging torque
waveforms for three different S
pp
values. The 1, 2/7 and 3/7 S
pp
combinations have
LCMs of 42, 84 and 126 respectively. Results verify the reduction in cogging torque
magnitude as the LCM increases.
Fig. 3.4 Cogging torque waveforms for various slot and pole combination
From the FE results shown in this section, it can be seen that the choice of slot and pole
configurations directly affect both the winding factor and the cogging torque
magnitudes. In the abovementioned comparison, the 18-slot, 14-pole model seems to be
-1.3%
-0.8%
-0.3%
0.3%
0.8%
1.3%
0 0.2 0.4 0.6 0.8 1
T
o
r
q
u
e

(
%

R
a
t
e
d
)
Normalized Time (P.U.)
42-slots, 14-poles
machine (1 Spp)
12-slots, 14-poles
machine (2/7 Spp)
18-slots, 14-poles
mechine (3/7 Spp)
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
64
the most appropriate choice as it produces a high winding factor (0.902) and very low
cogging torque magnitude (as it has a high LCM of 126). Thus this combination will be
used as the basis for the prototype design in this thesis.
Other suitable candidates for this study would be the 12-slot, 10-pole model as well as
the 18-slot, 16-pole model. The 18-slot, 14-pole model was chosen over the 12-slot, 10-
pole model as a lower cogging torque is desired. (The LCM produced by the 12-slot,
10-pole model the LCM was 60 as opposed to the 126 in the chosen configuration).
The 18-slot, 14-pole model was chosen over the 18-slot, 16-pole model because a
higher base speed was required. (An approximate base speed with 16-poles would be
375RPM as opposed to 429RPM in the 14-pole model).
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
65
3.3 PERFORMANCE IN COMPARISON TO DISTRIBUTED WINDING IPM
MACHINE
The growing popularity of CW is due to factors such as:
Shorter end-windingsdue to non-overlapping coils.
Less amount of copper used.
Simplified and possible automation of manufacturing process [105].
Higher slot-fill factor [104, 105].
Lower stator copper loss.
Higher tolerance to phase faults [20].
Additional leakage inductance increases CPSR in SPM machines [130].
There are, however, also advantageous factors which still make DW a preferred choice
over CW. Some of these are:
DW generates less MMF harmonics leading to lower core and magnet losses.
Higher saliency ratio leading to higher reluctance torque.
Unity winding factor can be achieved resulting in larger electro-dynamic torque.
High performance control techniques including sensorless control are readily
available.
In this section, the key differences between CW and DW will be discussed.
3.3.1 Airgap Flux Harmonics
The popularity of DWis due to its ability to produce MMF waveforms which are close
to sinusoidal. On the contrary, CW produces MMF waveforms which are rich in
harmonics. The MMF produced by stator coils can be expressed follows:
J = B
s
A
g
J
u

(3.7)
where,
B
s
=Flux density fromthe stator poles
A
g
=Airgap surface area
J
u
=Airgap reluctance
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
66
If the airgap length and overall slot opening widths are kept constant, the MMF is
directly proportional to the airgap flux density produced by the stator current. The
airgap flux density of a typical 14-pole, double-layer, DW machine with magnets
removedis shown in fig. 3.5:
Fig. 3.5 Flux waveform and corresponding frequency spectrum of a 14-pole, double-layer, DW
machine model
The only term that contributes to electro-dynamic torque production is the fundamental
component. The other terms contribute to increased core and magnet eddy current losses,
as well as additional leakage inductance. A detailed mathematical explanation of how
the additional harmonic components lead to increased frequency-related losses will be
shown in chapter 5.
Due to CW being non-sinusoidal in nature, it is expected that significant harmonic
components will be present in the airgap flux waveform. Airgap flux waveforms in
single- and double-layer, 18-slot, 14-pole CW models with magnets removed are shown
in fig. 3.6a and 3.6b respectively:
-0.25
0
0.25
0 100 200
Tesla
0
0.1
0.2
0.3
0 25 50
Tesla
Fundamental
component
(torque-producing
term)
00 00
Airgap circumference(mm)
25
Frequency components (n-order)
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
67
(a) 14-pole, 18-slot, single-layer CW machine model
(b) 14-pole, 18-slot, double-layer CW machine model
Fig. 3.6 Flux waveform due to armature reaction for single- and double-layer CW
Comparatively, the double-layer model produces much lower MMF harmonic
components than the single-layer model. Due to the nature of the machine designed for
this work, (where operating frequency can exceed several times the base frequency
making the machine more susceptible to frequency related losses), the double-layer CW
was selected as the base layout in this work.
-1
-0.5
0
0.5
1
0 100 200
Tesla
0
0.1
0.2
0.3
0 25 50
Tesla
-0.5
-0.25
0
0.25
0.5
0 100 200
Tesla
0
0.1
0.2
0.3
0 25 50
Tesla
Fundamental
component
(torque-producing
term)
Fundamental
component (torque-
producingterm)
00 00
Airgap circumference(mm) Frequency components (n-order)
Airgap circumference(mm) Frequency components (n-order)
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
68
3.3.2 Saliency Ratio and Constant Power Capability
The two main factors contributing to the field-weakening performance of IPM machines
are saliency ratio and characteristic current. Saliency ratio ( ) given by (3.8),
contributes to the additional reluctance torque as shown in (3.1), which is additive to the
total IPM machine torque.
=
I
q
I
d
(3.8)
where,
I
d
= d-axis inductance
I
q
= q-axis inductance
In an IPM machine I
d
< I
q
, resulting in > 1.
Measuring the saliency in integral-slot windings is less time consuming as the rotor can
be positioned according to the d- and q-axis flux paths (fig. 3.7) and the d- and q-axis
inductances can be measured accordingly. Fig. 3.7a shows the flux path across the pole
axis, while 3.7b shows the flux path over the inter-pole axis.
(a) Flux being channeled to the d-axis (pole
axis)
(b) Flux being channeled to the q-axis (inter-
pole axis)
Fig. 3.7 14-pole DW IPM with fluxbeing channeled to the d- and q-axis
In fractional-slot CW, the d- and q-axis flux paths are not obvious due to the
aperiodicity between slots and poles as shown in fig. 3.8.
q-axis
d-axis
q-axis
d-axis
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
69
Fig. 3.8 14-pole, 18-slot, CW-IPM flux plot showing no obvious d or q-axis flux paths
To measure d- and q-axis inductances in CW-IPM machines, the AC standstill test is
deemed most suitable [184]. The AC standstill test was implemented in FE analysis and
its accuracy was verified against the UNSW 4-pole segmented DW-IPM machine
(shown in Appendix A). With the FE results agreeing with the range of measured values,
this method was then used to measure the saliency ratio of the CW-IPM machine. To
account for saturation, measurements were taken at different current values. It should
also be noted that the effect of cross coupling was not studied as it would not
significantly affect the results.
With this method, inductances in the d and q-axes are determined from the self-
inductance (I
u
(0

)) and mutual-inductance (H
ub
(0

)) waveforms expressed by:


Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
70
I
u
(0

) =
_
[
I
u
I
s

2
-R
2
2n

(3.9)
H
ub
(0

) =
I
b
2n I
s

(3.10)
where,
I
u
=Voltage drop across the excited phase A,
I
b
=Voltage induced in the phase B due to excitation in phase A,
R =Stator resistance
=frequency of excitation.
0

=Rotor position
I
s
=Input current
By plotting I
u
, H
ub
versus rotor position and using curve fitting or FFT, the DC and
second harmonic terms of the respective waveforms can be determined. Subsequently,
the d- and q-axis inductances (I
d
and I
q
) can be determined by (3.11) and (3.12). The
derivation of these equations can be found in [184].
I
d
=
3
2
(I
dc
-H
dc
) -_
I
2
2
+H
2
]
(3.11)
I
q
=
3
2
(I
dc
-H
dc
) +_
I
2
2
+H
2
]
(3.12)
where,
I
dc
=DC term of the self-inductance
I
2
=Second harmonic term of the self-inductance (3 times base frequency for this 3-
phase machine)
H
dc
=DC term of the mutual-inductance
H
2
=Second harmonic term of the mutual-inductance (3 times base frequency for this
3-phase machine)
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
71
As a comparison, I
u
and H
ub
, for the UNSW Segmented DW-IPM is shown in fig. 3.9
below. This machine produced a saliency ratio of over 1.7.
Fig. 3.9Inductance waveform measured from UNSW Segmented IPM machine [174]
I
u
and H
ub
for CW-IPM machines generally have a lower second harmonic term and a
larger DC term, indicating lower saliencies. The inductance waveforms for an 18-slot,
14-pole model shown in fig. 3.10, resulted in a very small saliency ratio of 1.04.
Fig. 3.10Inductance waveform of an 18-slot, 14-pole CW-IPM machine
2.0E-03
2.2E-03
2.4E-03
2.6E-03
2.8E-03
3.0E-03
0 10 20 30 40 50
H
e
n
r
y

Deg
Self inductance (L
a
)
-8.0E-04
-7.0E-04
-6.0E-04
-5.0E-04
-4.0E-04
0 10 20 30 40 50
H
e
n
r
y

Deg
Mutual inductance (M
ab
)
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
72
A comparison of CW and DW with the same IPM rotor was done in [114]. The results
show that CW resulted in a lower saliency ratio. This was due to a significant increase
in the d-axis inductance, but with q-axis inductance remained almost the same as shown
in fig. 3.11:
(a) D-axis inductance comparison (b) Q-axis inductance comparison
Fig. 3.11 d- and q-axis inductance comparison [114]
As a consequence of a decrease in I
d
, the saliency ratio is decreased. Hence additional
reluctance torque contributions from IPM machines with CW are not as significant as
compared to DW-IPM machines. This implies that for CW-IPM machines, the CPSR
should be optimised by satisfying characteristic current equilibrium conditions rather
than achieving a high saliency ratio.
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
73
3.3.3 End Winding Length
The key advantage of CW is the reduction in end-winding length due to non-
overlapping coils. This enables the effective stack length of the machine to be increased
in a given amount of space. Fig.3.12compares end-winding lengths of DW and CW:
(a) Single-layer distributed
winding
(b) Single-layer concentrated
winding
(c) Double-layer concentrated
winding
Fig. 3.12 Estimated reduction in end winding length
Compared to single-layer DW (Fig. 3.12a), the end-winding length of a single-layer CW
(Fig. 3.12b) is nearly half, and for a double-layer CW (Fig. 3.12c), a quarter. For CW,
the end-winding length is also largely dependent on factors such as the quality as well
as the format of the winding.
x
x
2
x
4
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
74
3.3.4 Slot-fill Factor
The slot-fill factor otherwise known as the packing factor is inversely proportional to
the copper loss in the machine. Therefore, having the highest possibleslot-fill factor is
essential for achieving optimal efficiency in the machine.
The advantage that CW has over DW is that coils are wound around individual stator
teeth. This allows the use of more advanced winding methods such as the joint lapped
core method [105, 106], and prepressed windings in separable tooth pieces [104]. Fig
3.13a and 3.13b show the abovementioned processes respectively:
(a) J oint-lapped core windings [106] (b) Prepressed windings in a welded powered-
iron core stator [104]
Fig. 3.13 Advanced winding methods to achieve a high saliency ratio
Typical slot-fill factors of up to 35% can be achieved for DW and up to 45% can be
achieved for CW by traditional hand winding methods (based on our constructed IPM
machines). With the joint-lapped core and prepressed winding CW methods, a slot-fill
factor of up to a 78% can be achieved. Thus, these methods can effectively reduce
copper loss and end-winding length. Furthermore, with the joint-lapped core method,
the winding process can be automated, making the production process quicker.
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
75
3.4 COMPARING IPM AND SPM MACHINES WITH CONCENTRATED
WINDINGS
As shown in section 3.2, the saliency ratio of IPM machines with CW falls significantly,
compared to IPM machines with DW. This makes the electromagnetic performance of
the CW-IPM machine closer to that of a CW-SPM machine. Extensive studies on CW-
SPM machines have been carried out by El-Refaie and J ahns, with some of the key
contributions shown in [18, 90, 98, 116, 130, 131, 167, 185]. They managed to prove
that with fractional-slot CW, a very wide CPSR could be achieved in the SPM machine
a machine that was known to have very little or no field-weakening capabilities. Fig.
3.14 shows common forms of SPM and IPM rotors:
(a) Surface mounted
permanent magnets
(b) Interior single-piece/pole
permanent magnets
(c) Interior v-shaped
permanent magnets
Fig. 3.14 Performance comparison of different magnet shapes
Compared to the SPM machine, the IPM machine has the following advantages:
Buried magnets increase the mechanical robustness of the rotor
Magnets are not directly exposed to the airgap flux making them less prone to
large eddy current losses and risk of demagnetisation.
Greater flexibility in rotor variations, allowing flux concentrations and saliency
ratio alterations.
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
76
3.4.1 Airgap Flux Produced by the Magnets
This sub-section will compare airgap flux produced by the rotor permanent magnets
from fig. 3.14.
A general relation between magnet pole and airgap area can be described in (3.15).
B
g
=
A
m
A
g
/
(1 +A
m
R
g
)
B

(3.15)
where,
B
g
=Total airgap flux density due to the magnets
A
m
=Magnet pole area
A
g
=Airgap area
A
m
=Magnet internal permeance
R
g
=Airgap reluctance
B

=Magnet remanent flux density


The equation above tells us that a larger magnet pole surface and smaller airgap length
would result in greater airgap flux density. In the comparison done here, airgap length is
kept constant. Key model parameters are shown in table 3.3. This equation holds when
comparing two SPM machines with different pole surface areas. But when comparing
the SPM rotors with IPM rotors, (especially for IPM with v-shaped magnets), saturation
in the rotor core significantly affects the airgap flux. Fig. 3.15 shows saturation of the
inter-pole and pole areas of the rotor core:
Table 3.3
Key Specifications of Models with Different Magnet Shapes
Rated stator current 15A
rms
Total number of series turns per phase 133turns
Air gap length (IPM) 1mm
Air gap length (SPM) 1mm +Magnet depth
Magnet volume per pole 2870mm
3
Magnet remanent flux density 1.3T
Base speed 428.571rpm (50Hz)
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
77
Fig. 3.15 Flux density plot of the v-shaped IPM model showing saturation regions
Saturation in the core can be separated into two portions, the first being the saturation of
the inter-pole area, and the other, the saturation of the pole area. The saturation effects
in the inter-pole area are much stronger, thus the actual magnet span of the magnet is
reduced. The saturation effects in the pole sections are mild; this only slightly increases
the permeance of the magnet flux path through the rotor core to the airgap. (3.15) can be
modified by adding two constants to take saturation into account as follows:
B
g
=
k
p,spun
A
m
A
g
/
(1 +A
m
R
g
)
k
p,poIc
B


(3.16)
where,
k
p,spon
=permeability constant affecting actual span of the magnet pole
k
p,polc
=permeability constant reducing the flux density (from B
r
)
k
p,poIc
limits the peak airgap flux while k
p,spun
reduces the peak width (causing it to
fall off faster) as shown in fig. 3.16below.
Saturation of
inter-pole area
Saturation of
pole axis
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
78
(a) Airgap flux produced by SPM rotor (b) Airgap flux produced by IPM rotor with
single-piece/pole magnets
(c) Airgap flux produced by IPM rotor
with v-shaped magnets
Fig. 3.16Airgap flux produced bythedifferent rotor configurations
The SPM machine has a pole surface area of 1418mm
2
per pole, whereas the IPM with
single-piecemagnets has a pole surface area of 1327mm
2
per pole. Naturally, due to a
smaller pole surface and the inclusion of saturation effects in the latter design, the
overall airgap flux density is lower. The magnet surface area per pole for the v-shaped
IPM rotor is 2165mm
2
, significantly larger than the other two designs. However, due to
saturation effects, the flux density waveform (shown in 3.16c) is narrower and of
slightly lower magnitude compared to the SPM machine. When compared to the single-
piece IPM, the v-shaped IPM produced a larger peak value of flux density. But, due to
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
79
having higher saturation in the inter-pole area, the waveform is narrower resulting
only in a slightly higher RMS airgap flux density.
3.4.2 Constant Power Speed Range Comparison
Here, the CPSR and peak power performance of the three rotors in fig. 3.14 are
compared. These three rotors are modelled with the same CW stator with equal
excitation current. The rotor field is weakened by the variation of the current angle (y)
to maintain a constant power after base speed. In particular, the current angle is
increased as greater negative d-axis current is required to weaken the permanent magnet
field. This supresses the back EMF induced in the phase coils- hence helping to
maintain constant power over an increased speed range. It is assumed that the initial
rotor position is 0 (the rotor EMF axis is in line with the stator current axis).
I
u
= I
mux
sin(2nt +y) (3.17)
I
b
= I
mux
sin _2nt -
2n
3
+y]
(3.18)
I
c
= I
mux
sin _2nt +
2n
3
+y]
(3.19)
Fig. 3.17Peak power and CPSR comparison between three rotor types
0
200
400
600
800
1000
1200
1400
0 500 1000 1500 2000 2500 3000 3500
Surface Magnets (1mm+AG) Flat-Shaped Magnets (1mmAG)
V-Shaped Magnets (1mmAG)
Speed (RPM)
P
o
w
e
r

(
W
)

e
base
4.3:1
CPSR
5.6:1
CPSR
7.2:1
CPSR
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
80
Fig. 3.17shows that the SPM machine had the highest power density from base speed
up to the 4:1 CPSR point. The IPM machine with v-shaped magnets, on the other hand
produced slightly less power but a wider CPSR. The IPM machine with rectangular
single-piece/pole magnets was not able to outperform the other two designs in terms of
both power density and CPSR. The results indicate that the IPM machine with v-shaped
magnets would be the most suitable candidate to achieve good field-weakening
characteristics.
Chapter 3: Investigation of the Concentrated Winding IPM Machine for Wide Field-weakening
Applications
81
3.5 CONCLUSION
Chapter 3 has investigated the suitability of implementing CW on the IPM machine. It
was shown that due to the fractional-slot distribution used in CW, sinusoidal back EMF
and low cogging torque can be achieved at a small expense of a lower winding factor
compared to equivalent integral-slot machines. By comparing several 14-pole layouts, it
was seen that the 18-slot model produced a sinusoidal back EMF waveform with the
highest magnitude of induced voltage of the models.
In a comparison of MMF waveforms, it was shown that the MMF waveforms produced
by CW contain much higher harmonics compared to DW models. Additionally, it was
shown that single-layer CW contains significantly higher harmonic content compared to
double-layer CW.
Different winding methods were investigated and it was shown that by advanced
winding methods, a slot-fill factor of up to 78% can be achieved.
Lastly, this chapter showed that the CW-IPM machine with rectangular single-
piece/polemagnets is inferior to the CW-SPM machine in terms of torque density and
field weakening capability. The CW-IPM with v-shaped magnets, on the other hand,
produced improved field weakening capability, but with slightly lower peak torque
density as the CW-SPM machine due to saturation effects. It should also be noted that
since fractional-slot windings result in reduced saliency ratio in IPM machines, the
same results will be expected if fractional-slot DW are used that is, lower torque
density is expected in IPM compared to SPM machines.
The next chapter will build upon the findings in this chapter to arrive at a final
optimised design.
82
CHAPTER 4
DESIGN OF AN IPM MACHINE WITH CONCENTRATED
WINDINGS FOR FIELD WEAKENING APPLICATIONS
4.1 INTRODUCTION
General design goals for PM machines used for field-weakening applications include:
Achieving a very wide CPSR
High efficiency in the constant power region
High torque densities with low torque ripple
In this work, the CW-IPM will be compared with two equally-sized DW-IPM machines.
To demonstrate that the CW-IPM is a suitable candidate for field-weakening
applications the following specific goals are set:
Achieving sinusoidal back EMF waveform with high winding factor
Low cogging torque
High torque density With a shorter end-winding length, the effective stack
length of the machine can be increased. The output power of this design should
exceed 580W, which has been achieved by both UNSW IPM machines.
A very wide CPSR greater than 4:1
High efficiency exceeding 85% in the constant power region.
In this chapter, the CW-IPM is designed and optimised in order to achieve these design
goals. The method to optimise the CPSR by achieving characteristic current equilibrium
is shown; rotor geometry and airgap length are also varied to further increase the CPSR.
Material considerations and manufacturing methods are looked into. Mechanical
stresses on the inter-pole link sections of the rotor are also investigated. Lastly, full
specifications of the final FE design will be stated and predicted performance
characteristicswill be presented.
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
83
4.2 CONDITIONS FOR MAXIMISING THE CPSR
As mentioned in chapter 3 the saliency ratio in a CW-IPM machine is reduced due to
the significant increase in d-axis inductance and with q-axis inductance remaining
relatively constant.
In order to achieve a high saliency ratio, the inter-pole or q-axis flux paths should have
as little resistance as possible, while the pole or d-axis should have barriers to create
resistance. Based on these principles, axially laminated (fig.4.1a) and multi-barrier
(fig.4.1b) designs are often used to achieve high saliency ratios [12, 186].
(a) 4-pole axially laminated rotor (b) 4-pole multi-barrier rotor
Fig. 4.1 Rotor types used for increasing saliency ratio [12, 186]
Flux paths in integral-slot DW machines are periodic, thus the rotor pole geometry can
be shaped to channel d- and q-axis flux. However in CW machines these flux paths are
not obvious, (as shown in fig. 3.8). Attempts to increase the saliency ratio by rotor
magnet geometry variations (appendix B), showed that saliency ratio increments with
optimised designs were small the maximum saliency ratio achieved was 1.14. Most of
the designs achieving higher saliencies were also impractical as it required very thin
inter-pole link sections, which would be subjected to high mechanical stresses during
high speed operation.
q-axis
flux path
d-axis
flux path
q-axis
flux path
d-axis
flux path
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
84
Additionally, in [187, 188], the saliency ratios for double-layer CW as well as for
different S
pp
combinations were also investigated, (this comparison is shown in
appendix C). In this comparison, it was also shown that the saliency ratio remained
relatively low, (between 1.06 and 1.11).
Thus the rotor magnet geometry was selected based on its ability to produce the
required magnet flux-linkage and to maintain mechanical robustness of the rotor, rather
than to achieve the optimal saliency ratio.
With a low saliency ratio achieved in the CW-IPM machine, the key condition for
optimising CPSR is achieved by satisfying the characteristic current equilibrium
conditions as shown in (4.1).
I

= I
`
ch
=
E
`
0(I-n)

c0
I
d
=

pm
I
d

(4.1)
where,
I

=Rated current

pm
=Permanent magnet flux linkage
The following shows a breakdown and the relationship between these two terms (
pm
and I
d
).
I
d
is affected by both the self-and mutual inductance in the following relationship:
I
d
=
3
2
(I
dc
-H
dc
) -_
I
2
2
+H
2
]
(4.2)
where,
I
dc
=1
st
harmonic component of self-inductance
H
dc
=1
st
harmonic component of mutual-inductance
I
2
=2
nd
harmonic component of self-inductance and
H
2
=2
nd
harmonic component of mutual-inductance
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
85
Self-inductance I
u
and mutual-inductances H
ub
vary as a function of the armature
windings as well as effective airgap. They can be expressed in terms of machine
parameters, shown in (4.3) and (4.4) respectively [130].

I
u
(0) =
p
0
r
ug
l
c]]
l
g
(0)
_ N
u
2
(0)
2n
0
J0 (4.3)
H
ub
(0) =
p
0
r
ug
l
c]]
l
g
(0)
_ N
u
(0)N
b
(0)
2n
0
J0
(4.4)
where,
p
0
=Permeability of free space
r
ug
=Airgap radius
l
c]]
=Stack length of the machine
l
g
=Airgap length
N
u
=Phase A winding function
N
b
=Phase B winding function
In CW machines, the magnitude of self-inductance is high, while the magnitude of
mutual-inductance is relatively low compared to DW machines [20]. Therefore the
magnitude of I
d
varies essentially with self-inductance, and the following conditions
can be assumed:
I
d
l
c]]
I
d
r
og
I
d
N
o
2
The variation of these parameters may also affect
pm
, which can be generally
described in terms of the number of series turns under a pole (N
s
) exposed to airgap
flux (
ug
) as follows [3, 174, 189]:

pm
= N
s

ug

(4.5)
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
86
where,
ug
is made up of two terms: the flux density normal to the pole surface (B
s
) and
the pole area (A
p
). (4.5) can be expanded further as follows:

pm
= N
s
B
s
A
p
=N
s
B
s
r
o
0
p
l
c]]
(4.6)
where,
r
o
=Rotor outer radius(proportional to airgap radius r
ug
in (4.3))
0
p
=Pole span
B
s
can further broken down to give:
B
s
=
B

1 +
p
(mug)
l
g

m

(4.7)
where,
p
r(mog)
=Magnet relative permeability

m
=Magnet thickness
Substituting (4.7) into (4.6), we get:

pm
=
B

N
s
r
o
0
p
l
c]]
1 +
p
(mug)
l
g

m

(4.8)
and the following relation can be attained:

pm

-1
l
g

pm
l
c]]

pm
r
ro
r
og

pm
N
s

pm
0
p

pm

m

pm
B


Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
87
The relation between the abovementioned variables relating to
pm
and I
d
indicate that,
in order to effectively change the characteristic current, three parameters, (independent
of I
d
), can be varied. They are: 0
p
,
m
and B

.
Taking B

as example to see how it can be used to satisfy characteristic current


equilibrium conditions, we shall assume the machine geometry and number of turns are
kept constant (thus L
d
is constant). Then, with a range of rated current values, B

can be
chosen to satisfy characteristic current equilibrium conditions and achieve optimal
CPSR, as shown in the contour plot (fig. 4.2):
Fig. 4.2Contour plot showing the variation of magnet remanent flux density versus rated
current
Fig. 4.2 shows that with an almost linear relation between the rated current and magnet
remanent flux density (Green region), the CPSR up to a range of 10 to 15:1 can be
achieved if equilibrium conditions are met. The steep variations in contours indicate the
importance on achieving characteristic equilibrium conditions in the optimisation of
CPSR. The plot also shows that the optimal region narrows towards the end as the steel
gets saturated with higher currents.
1.09T
1.13T
1.18T
1.24T
1.29T
3A 3.25A 3.5A 3.75A 4A 4.25A 4.5A 4.75A 5A 5.25A 5.5A
10-15 5-10 0-5
(RatedCurrent)
(
M
a
g
n
e
t

R
e
m
a
n
e
n
t
F
l
u
x

D
e
n
s
i
t
y
)
CPSR (X : 1)
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
88
4.3 CHOICE OF WINDING ARRANGEMENT
The choice of winding arrangement affects the winding factor, which in turn affects the
efficiency, output power and field-weakening performance of the machine. Unlike
integral-slot DW, the arrangement of phase coils in fractional-slot CW to achieve the
desired winding factor is not as straight forward and is often hard to visualise. Generally
the windings should be arranged such that maximum flux is linked from the magnets to
the stator coils. A method for obtaining the optimal phase coils arrangement in CW to
achieve the highest possible flux linkagewas illustrated by J . Cros and P. Viarouge in
[16].
The following example briefly explains this method by the use of an arbitrary 12-slot,
14-pole, double-layer model (with S
pp
=2/7).
Step 1: Using the number of slots per pole per phase (S
pp
), a sequence of 1s and 0s
can be obtained. The numerator represents the number of 1s and the denominator
represents the total number of integers in the sequence.
For S
pp
=
2
7
, there are two 1s for a sequence of seven integers with the rest being 0s.
1 1 0 0 0 0 0
Step 2: The sequence is re-written such that the 1s and 0s alternate with the ones
being the first integer.
1 0 1 0 0 00
Step 3: Repeat this sequence until the total number of 1s equal to the number of slots
(in this case with twelve slots, the sequence repeats six times).
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
89
Step 4: Under each integer the winding sequence A C B A C B A CB A C B . . . . .
is writtenout:
Step 5: The integers under 1s will be used as one layer of the winding
Step 6: The other layer will be filled with the opposite polarity in the neighbouring slot.
With this winding method, flux-linkage will be maximised and the winding factors for
various S
pp
values as shown in table 3.1 can be achieved.
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
90
4.4 MATERIAL CONSIDERATIONS
Selection of materials depends on the requirements of torque density, weight, cost,
operating temperature range and external demagnetisation fields. The three key
materials to consider are: permanent magnets, core and coil material [190]. This section
explains how the materials were selected for this design, covering key properties such as
saturation, energy density and thermal considerations. Losses of these materials will be
discussed in greater detail in chapter 5.
4.4.1 Permanent Magnet Material
Permanent magnet characteristics can be described by the demagnetisation curve shown
in fig. 4.3. From this curve desired magnet parameters such as - remanent flux density
(B
r
), coercivity (H
c
), and recoil permeability (
r
) can be determined.
0
Fig. 4.3. Permanent magnet Demagnetisation curve
There are several types of permanent magnet materials in the market. Ferrite (FeO),
Alnico (AlNiFeCo), Samarium-cobalt (SmCo) and Neodymium-iron-boron (NdFeB)
are more commonly used. The advantages and disadvantages of these materials are
stated in table 4.1 below [191-194].
0
Normal
Curve
H H
B
H
I
n
c
r
e
a
s
i
n
g

T
e
m
p
e
r
a
t
u
r
e B
r
iH
c

r
=
dB
dH
Fig 43
HH
c
HH H
c
B
Intrinsic
Curve
Demag.
Paths
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
91
Table 4.1
Advantages and disadvantages of different magnet types
Magnet type Advantages Disadvantages
Ferrite Least expensive magnet material
High operating temperature up to
300 C
Hard and brittle
Lowest remanent flux
density of up to 0.42T
Alnico High operating temperature up to
520 C
Lowest temperature coefficient
(0.02%/C)
Extremely hard and brittle
Can be easily
demagnetised
Samarium-cobalt High remanent flux density of up to
1.16T
Extremely resistant to corrosion
High resistance to demagnetisation
High operating temperature of up
to 350C
Low temperature coefficient
(0.04%/C)
High coercivity
Extremely hard and brittle
Most expensive magnetic
material
Neodymium-
iron-boron
Highest remanent flux density of
up to 1.48T
High resistance to demagnetisation
Least brittle
Lower cost compared to SmCo
High coercivity
Low operating temperature
up to 200C
High temperature
coefficient (0.12%/C)
Susceptible to corrosion
NdFeB magnets are the preferred choice in present day applications due to its
reasonable cost, high coercivity and high remanent flux density [195].
However, NdFeB magnets have comparatively low operating temperatures, which cause
the knee from the third quadrant to enter the second quadrant if the temperature gets too
high (fig. 4.5). When the knee enters the second quadrant, (the coercivity of the material
decreasing), and with armature flux opposing the flux from the magnets, the risk of
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
92
irreversible demagnetisation is increased dramatically [3]. Therefore, the thermal
aspects of the magnets have to be considered in the design stage. To estimate
temperatures within the machine a basic 1kW CW-IPM model was created in Motor-
CAD (as shown in appendix D). The model showed that under worst-case scenario
conditions, the magnets would operate to under 110C. NdFeB magnets have maximum
operating temperatures of between 80C to 220C and curie temperatures of between
310C and 380C for different grades. The maximum energy product of the magnetic
grade is inversely proportional to the temperature rating of the magnet. For example, a
magnet grade with 80C operating temperature has remanent flux density of up to 1.48T,
whereas a grade with 220C operating temperature has a flux density of only up to 1.2T
[196]. The chosen grade of NdFeB magnet is N28UH with remanent flux density of
1.04 to 1.09T and operating temperature of 200C. Each magnet piece has a dimension
of 2mm x 13.5mm x 79mm and had nickel and copper coating (Ni-Cu-Ni) to prevent
corrosion.
4.4.2 Core Material
Silicon sheet steel is widely used in stator and rotor cores of electric machines. This is
due to properties such as: high saturation magnetisation, low cost and relatively low
losses [197, 198].
The silicon steel grade was selected based on the torque required and the tolerance to
losses. Saturation and loss are inversely proportional to each other. For example, grade
35J N210 has a saturation point (B
50
) at 1.65T and core loss of 2.05W/kg at 1.5T/50Hz,
whereas grade 50J N1300 has a saturation point (B
50
) at 1.76T andcore loss of 8.1W/kg
at 1.5T/50Hz.
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
93
For machines designed for field-weakening applications, cores with low losses are
preferred, as excitation frequency could exceed several times the base frequency.
Furthermore with CW, the additional high frequency leakage harmonics would result in
additional core losses. These issues will be studied in greater detail in chapter 5.
The chosen rotor and stator core material, (grade 35RM300 from Australian supplier
Sankey), is non-oriented silicon sheet steel with a thickness of 0.35mm. It has a
saturation point (B
50
) at 1.68T, and core losses of 2.6W/kg at 1.5T/50Hz. The
magnetisation and core loss curves provided by the supplier are shown in fig. 4.4 and
4.5 respectively. For the model, the declaration of material properties was based on a
saturation point (B50), at 1.65T with a 10% safety factor.
Fig. 4.4Magnetisation curve of 35RM300 (provided bySankey)
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
94
Fig. 4.5Core loss curve of 35RM300 (provided by Sankey)
4.4.3 Stator Coil and Insulating Material
The ideal coil material is one with infinite conductivity and the lowest temperature
coefficient of resistance [199]. Copper is a common choice for coil material due to its
relatively lower price, high conductivity, good mechanical strength and a relatively low
temperature coefficient of resistance. The amount of I
2
R loss per coil can be calculated
by (4.10).
I
2
R = 2N
coI
I
2
p
l
c]]
A
w

(4.10)
where,
N
coI
=Number of turns per coil
p =Conductor resistivity
A
w
=Cross-sectional area of a wire
l
c]]
=Machine stack length
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
95
The available cross-sectional area per wire is found based on the available stator slot
area (A
s
), slot-fill factor S
ff
and the number of turns per coil, as shown in (4.11). The
factor of 2 in the denominator is due to the chosen double-layer windings layout.
A
w
=
A
s
S
]]
2N
coI

(4.11)
Excessive amounts of I
2
R losses may increase the operating temperature causing
insulation degradation or even breakdown. This can be prevented by predicting losses
with temperature increases. Resistance of a conductor due to temperature rise can be
estimated by (4.12). The resistance and copper loss per phase in the final model as
temperature increases is shown in fig. 4.6below.
R = R
0
[1 +o(I -I
0
)] (4.12)
where,
I =Current operating temperature
I
0
=Initial operating temperature
R =Resistance of conductor at current operating temperature
R
0
=Resistance of conductor at initiation operating temperature
o =Temperature coefficient of resistivity (0.004041 for copper)
Fig. 4.6 Resistance and copper loss per phase as temperature increases
Present-day winding insulation materials such as Nomex

have operating temperatures


of up to 220C, and are able to withstand temperatures of up to 400C for several hours
based on Arrhenius law [200].
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
96
4.5 OPTIMISATION OF MACHINE GEOMETRY
One of the aims of this design was to achieve a high torque density with a specific size.
More specifically- the prototype machine must fit into the same ABB casing shown in
fig. 4.7which was originally used to house a 550W induction machine.
Fig. 4.7. ABB casing inserted 18-slot stator core
4.5.1 Stack Length
The stack length of the machine was determined based on the axial length of the
original ABB machine 55mm stack length and 27mm end-winding length per side, (fig.
4.8a). A plastic stator was built to estimate the end-winding length of a double-layer
CW stator. The horizontal-fill method (fig. 4.9a) was used and the end-winding length
was measured to be 14.5mm for a 40% S
ff
(fig. 4.8b).
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
97
(a) End-windings for double-layer DW stator (b) End-windings for plastic double-layer CW
stator
Fig. 4.8 End winding length comparison between UNSW DW stator with windings done on a
plastic model
Slots in double-layer CW can also be wound using thevertical-fill method, (fig 4.9b).
The advantage of the vertical-fill method is that the end-winding length will be
approximately half the end-winding length when wound with the horizontal method.
This results in lower copper loss, and also allows space for a longer stack. However,
this method is extremely time-consuming, (approximately 45 to 60 minutes for 90 coils).
Furthermore, this method increases the chances of having unequal number of turns per
phase as a former and counter can be used to pre-wind coils.
Although the horizontal-fill method results in a thicker end-winding length and higher
copper loss, this method is a lot less time consuming (20 minutes per coil with a
minimum slot opening with of 1.2mm) and a slot-fill factor of up to 44% can be
achieved. The likelihood of having unequal turns per coil is also lower, as a machine
can be used to aid the winding process.
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
98
(a) Horizontal-fill method
(b) Vertical-fill method
Fig. 4.9Hand winding methods
Despite the expected decrease in efficiency and power density, the horizontal method
was chosen to reduce the production hours, cost of production, and possibility of error
in this initial prototype.
Compared to the end-winding length of the DW machine (25.5mm per side), the end-
winding length of the prototype machine was 12.5mm per side, (shown in fig. 4.9).
Therefore, with an additional 26mmof axial space in the casing, the effective stack
length can be increased from 55mm to 80mm. This will allow the original power
density of the DW-machine (550W) to be increased to (550 X 80/55 =800W).
x
2
Side view Front view
x
Side view
Front view
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
99
4.5.2 Rotor Outer Diameter and Airgap Length
The rotor outer diameter was chosen to be 80mm, 1mm smaller than the two other
comparison DW-IPM machines. This was done to make space for a larger airgap.
Typically, motors of this size, (under 1kW), have airgap lengths between 0.3mm to
0.5mm. In CW motors, the additional leakage harmonic terms created result in localised
saturation and unbalanced pull forces [20]. This would cause increased core losses and
significantly reduce the field-weakening performance of the machine. To reduce the
effects of the leakage harmonic terms caused by armature reaction, the airgap length of
the IPM machine can be increased. It was shown in [201] that a larger airgap resulted in
a significant increase in CPSR. Conversely, it would also result in larger reluctance for
the permanent magnet flux linked with the stator coils. This would cause the output
power to decrease. Fig. 4.10 shows the effects of airgap length variation with output
power and CPSR.
(a) Airgap length variation with output power (b) Airgap length variation with CPSR
Fig. 4.10 Airgap length variation with output power and CPSR
The final airgap length chosen is was 1.2mm. This helps in achieving a wide CPSR
(>7:1) and an output power of approximately 800W. In order to sustain such high power
with such a large airgap, high energy magnets with maximised surface area per pole
500
700
900
1100
1300
1500
1700
1900
0 0.5 1 1.5
P
o
w
e
r

(
W
)
Airgap Length (mm)
4
5
6
7
8
9
10
0 0.5 1 1.5
C
P
S
R

(
X

:

1
)
Airgap Length (mm)
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
100
must be used. The maximisation of magnet surface resulted in rotor structural issues
with single-piece/pole, rectangular-shaped magnet designs, thus v-shaped magnets with
iron bridges were preferred. This will be discussed later in section 4.6 of this chapter.
4.5.3 Stator Geometry
The stator inner diameter of 82.4mm was obtained from the rotor outer diameter and
airgap length. The basic stator geometry can be defined by four main parameters: tooth
width, yoke length, shoe thickness, and slot opening width, as shown in fig. 4.11.
Ideally, small tooth width, yoke length and shoe thickness are desired. As such, a larger
slot area can be achieved to contain larger conductors, which in turn lowers copper loss.
A wide slot opening width (w
so
) is also desired for simplifying the winding process.
However, smaller areas in the teeth and yoke mean a higher saturation level in these
areas if large currents are used.
Fig. 4.11. Key parameters defining the stator geometry
Slot opening
width (w
so
)
Tooth width
(w
t
)
Yoke length
(l

)
Slot area
(A
s
)
Shoe thickness(
sh
)
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
101
Manually calculating the exact stator geometry to control saturation levels can be
complex. Thus a simple two-step process is used here Firstly, a linear relationship
between the flux and stator geometries is used as an initial estimation [202].
Subsequently FE analysis will be used to fine-tune the geometry to ensure that
equivalent flux is achieved in the tooth and back iron as well as to prevent saturation
from occurring at key portions of the stator.
From [202], the estimated yoke length and tooth width is given as follows:
The yoke length or back iron (l

) is given by:
l

=
n
s
B
ug
2B
okc
p
(4.13)
and the tooth width (w
t
) is given by:
w
t
=
n
s
B
ug
B
tooth
n
s
(4.14)
where,

s
=Stator inner diameter
B
ug
=Peak airgap flux density
p =Number of pole pairs
B
okc
=Yoke flux density
B
tooth
=Tooth flux density
n
s
=Number of stator slots
The desired core material has a saturation magnetisation of 1.68T, hence, allowing for a
10% safety margin, peak flux density of of the yoke and tooth was chosen to be 1.5T.
With the chosen peak flux density in the yoke and tooth, the estimated yoke length and
tooth width from (4.13) and (4.14) are 5.3mm and6.8mm respectively. As a first step of
optimisation, the stator was designed with these values. The design was then further
optimised by the use of FE analysis where non-linearities in the core material were
taken into account.
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
102
Fig. 4.12Flux density plots showing peak flux density in the yoke and tooth
The final yoke length and tooth width were 5.1mm and 7.4mm respectively. These
dimensions produced peak flux of 1.48T in the yoke and 1.47T in the toothas shown in
fig. 4.12.
Slot Opening Width
The FE results show that wider slot openings and smaller shoe thickness result in a
lower CPSR. However, smaller slot openings would lead to increased difficulty in
inserting the stator coils, and thicker shoes would lead to smaller slot areas. For this
design, theideal shoe thickness was found1.4mm (as small as possible to allow larger
slots, but big enough to prevent oversaturation of the shoe region) and a suitable slot-
opening width was determined as follows:
Color ShadeResults
Quantity: |Fluxdensity| Tesla
Time(s.) : 0.001Pos(deg): 2.571
Scale/ Color
13.07917E-6 / 187.43178E-3
187.43178E-3 / 374.85045E-3
374.85045E-3 / 562.26915E-3
562.26915E-3 / 749.68785E-3
749.68785E-3 / 937.10661E-3
937.10661E-3 / 1.12453
1.12453 / 1.31194
1.31194 / 1.49936
1.49936 / 1.68678
1.68678 / 1.8742
1.8742 / 2.06162
2.06162 / 2.24904
2.24904 / 2.43646
2.43646 / 2.62387
2.62387 / 2.81129
2.81129 / 2.99871
B =1.47T
B =1.48T
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
103
With the main considerations being the ease and time taken to wind the stator; at the
same time investigating the possible slot-fill factor, plastic stator portions were made
with three different slot opening widths: 1.2mm, 1.6mm and 2mm as shown in fig. 4.13.
Fig. 4.13Plastic stator made with three different slot opening widths
60 turns of winding, diameter 0.80mm, were fitted into the respective slots. The
duration taken to fit the conductors into the slots with respective slot openings were: 1.2
mm 15 mins; 1.6mm 12mins, and 2mm 5mins. The ideal (lossless) CPSR obtained
for the various slot opening widths from FE analysis was: 1.2mm 14:1; 1.6mm 12:1
and 2mm 9:1. Although more time was required to wind the stator asthe slot-opening
widths decrease, the slot-fill factor was the same with all three widths. Since the same
slot-fill factor could be achieved and with the preference of achieving a wider CPSR,
the 1.2mm was selected in the final prototype.
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
104
4.6 ROTOR DESIGN AND STRUCTURAL CONSIDERATIONS
Compared to the SPM machine, the IPM machine has buried magnets. The
disadvantage of having buried magnets is that the width and thickness of the magnets
are limited by the inter-pole link sections, especially if single piece/pole magnets are
used, (fig. 4.14a).
(a) single-piece/polemagnet design (a) v-shaped magnet design
Fig. 4.14 IPM rotors showing inter-pole link sections
One way to increase the torque density of the machine is touse v-shapedmagnets, (fig.
4.14b), which increase the surface area of the magnets and provides flux concentration.
It was shown in chapter 3 that v-shaped magnets produced higher torque density
compared to rectangular, single piece/pole magnets with the same volume. Additionally
v-shaped magnets provide additional flexibility of v-angle variation to achieve desired
airgap flux density. Figure 4.15 shows the variation of magnet surface area and
corresponding airgap flux density when the v-angle is varied:
Inter-pole
link
sections
Inter-pole
link
sections
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
105
(a) 120 v-angle (b) 60 v-angle
Fig. 4.15. V-angle variation
In order to maintain the characteristic current equilibrium, rated current had to vary with
changes to the v-angle. The power ratings and optimised CPSR with various v-angles
are shown in Table 4.2 below [201]. Magnet volume is not constant for this set of
results. The per unit power versus speed curves for the various designs are shown in fig.
4.16.
Table 4.2
Power and CPSR with variation of v-angle
v-angle Power at base speed CPSR
120 600W 4.2:1
90 725W 6:1
60 1050W 10:1
30 1835W 8:1
Fig. 4.16Normalised power versus speed characteristics withvariation of v-angle
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 1000 2000 3000 4000
120 degrees 90 degrees 60 degrees 30 degrees
Speed (RPM)
P
o
w
e
r

(
P
.
U
.
)
B
air
=0.70T
A
m
=587mm
2
v-angle =120
B
air
=0.97T
A
m
=951mm
2
v-angle =60
6:1
CPSR
8:1
CPSR
e
base
4.2:1
CPSR
10:1
CPSR
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
106
The above figure and table show that power increases as v-angle is reduced; however
CPSR starts to decrease past 60. Thus, an angle of 60 is chosen as the optimal value
for this particular design.
4.6.1 Structural Considerations
In an SPM rotor, the permanent magnets (which are inherently brittle), attached on the
outside of the rotor are structurally vulnerable [203]. Various methods are used to
maintain the mechanical robustness, but each presents its own problem. Glued magnet
pieces, (fig 4.17a), are the easiest to manufacture but, are not suitable for high speed
operations due to the possibility of magnets detaching fromthe rotor. Full-cylindrical
rotor magnets, (fig 4.17b), are more secure but are vulnerable to vibration and
expansion of the rotor core. Retaining sleeves, (fig 4.17c), are suitable for high speed
operations but present additional rotor losses.
(a) SPM rotor with Glued
magent pieces [204]
(b) SPM rotor with full-
cylindrical rotor magnets
[205]
(b) SPM rotor with magnets
enclosed in a retaining sleeve
[206]
Fig.4.17 Various types of SPM rotors
IPM rotors on the other hand, havethe advantage of magnets being contained inside the
rotor. (Fig. 4.18 shows an IPM rotor with buried single-piece/pole magnets). This
simplifies the rotor manufacturing process, reduces rotor magnet losses, and enhances
the mechanical robustness of the rotor.
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
107
Fig. 4.18 IPM rotor with buried single-piece/pole magnets
The key disadvantages of IPM rotors are: rotor core losses will be higher and the
magnet pole span and thickness are limited by the inter-pole link sections. The inter-
pole link section has to be made thinner and longer as the required magnet volume
increases, thereby subjecting them to large mechanical stresses, especially at high
speeds.
During operation, two types of forces are inflicted upon the rotor poles magnetic
forces from the stator coils, and centrifugal forces during high rotational speeds. The
surface force or magnetic pressure (P
m
) is expressed in (4.15) where the centrifugal
force (F
ccn
) is calculated based on (4.16), both of whichare acting in an outward normal
direction to the rotor surface [207].
P
m
= _
B
g
2p
0 s
Js
(4.15)
F
ccn
=
m(r
m
)
2
r
r
(4.16)
where,
p
0
=Permeability of free space
m =Mass of object subjected to force
r =Radius of cylinder
Buried magnets
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
108
To obtain a general estimate of the magnetic pressure created by stator and rotor pole
interaction, a basic solenoid-magnet model, (as shown in fig. 4.19), was created. It
simulates the worst-case over-current scenario, where the airgap flux density was
modelled to be 1.5T. The airgap length was modelled as 0.6mm to account for any rotor
eccentricities.
Fig. 4.19Modelled solenoid-magnet model
The resultant attractive force acting in the direction normal to the magnet surface is
1075N. This would be equivalent to approximately 1MPa of pressure acting normal to
the magnet surface.
-1
-0.5
0
0.5
1
1.5
0 5 10 15
mm
Tesla
Regionvectorsresults
Quantity:FluxdensityTesla
Colour scale
26.35732E-3 / 309.69441E-3
309.69441E-3 / 593.03149E-3
593.03149E-3 / 876.36858E-3
876.36858E-3 / 1.15971
1.15971 / 1.44304
1.44304 / 1.72638
1.72638 / 2.00972
2.00972 / 2.29305
2.29305 / 2.57639
2.57639 / 2.85973
2.85973 / 3.14307
3.14307 / 3.4264
3.4264 / 3.70974
3.70974 / 3.99308
3.99308 / 4.27641
4.27641 / 4.55975
Tooth Coil Coil
Magnet
0.6mm Airgap
Airgap
Flux
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
109
To calculate centrifugal force, the v-shaped magnet model was used. It included the
mass of the magnets as well as the rotor steel.
Fig. 4.20Sections for centrifugal force calculation
Due to a relatively small rotor radius (40mm), the calculated centrifugal pressure at
6000rpm (maximum speed) was small (174kPa). Adding the centrifugal pressure to the
pressure resulting from the magnetic forces, total outward normal pressure of 1.174MPa,
acting on each pole section, was attained. In order to analyze the effects of these forces
on the inter-pole link sections, an FE model as shown in fig. 4.21, was created.
Figure 4.21Model showing outward normal pressure on each section of the rotor
The mechanical data for the selectedsteel grade is given in table 4.3.
Silicon steel section
Density =7600 kg/m
3
Volume =7.73X10
-6
m
3
Mass =58.71g
Magnet sections
Density =7500 kg/m
3
Volume 2.025 X 10
-6
m
3
Mass =15.19 g/piece
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
110
Table 4.3
Mechanical data for silicon steel grade (J FE-35J N210) [198]
Youngs modulus 1.7e
5
[MPa]
Poisson ratio 0.28
Density 7.6e
-6
[kg/mm2]
Thermal expansion 1.2e
-5
[1/C]
Tensile yield strength 400 [MPa]
Compressive yield strength
#
400 [MPa]
Ultimate tensile strength 515 [MPa]
#
Estimated quantity Data not found
Three different models with 0.7mm inter-pole link sections were compared-
Rectangular single-piece/polemagnet model, v-shaped magnet model (60 v-angle), v-
shaped magnet model with 0.7mm iron bridge. Maximum principal stress points of
these three models were shown in Fig. 4.22a to 4.22c, respectively.
(a) single-piece/polemagnets (b) v-shapedmagnets no bridges
(c) v-shaped magnets with bridges
Fig. 4.22 Mechanical stress analysis of rotor
15.3MPa
12.5MPa
25.1MPa
133.5MPa
103.6MPa
154.6MPa
130.3MPa
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
111
The mechanical model shows that the maximum stress imposed on all the three models
was within the mechanical limits of the chosen silicon steel grade. However, the model
with v-shaped magnet and bridges was preferred, as it allowed greater flexibility in
minor alterations to the magnet slot size andv-angle to eliminate any unforseen errors.
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
112
4.7 FINAL MANUFACTURED DESIGN
The final CW-IPM design was a 14-pole, 18-slot, double-layer winding layout with v-
shaped magnets. The final design was chosen based on the appropriate selection of
slot/pole combination and magnet shape (shown in chapter 3); as well as the
optimisation and material selection process shown earlier in this chapter. Here, key
machine parameters and torque performance characteristics of the final FE design will
be illustrated. Full specifications of the final design are shown in Table 4.4, and
technical drawings are shown in Appendix C.
The final winding layout is shown in fig. 4.23 (PA, PB and PC represents positive
terminals for each phase coil while MA, MB and MC represent negative terminals).
Fig. 4.23Final 18-slot, double-layer winding layout
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
113
Table 4.4
Specifications of thefinal design
Stator outer diameter 130mm
Rotor outer diameter 80mm
Airgap length 1.2mm
Stator inner diameter 82.4mm
Shaft outer diameter 24mm
Stack length (stator) 80mm
Stack length (rotor) 79mm
Area per half slot 86.3mm
2
Stator and Rotor lamination thickness 0.35mm
Slot-opening width 1.2mm
Number of slots 18slots / double-layer
Winding type Double-layer concentrated
Packing (slot-fill) factor 41.5%
Number of turns per coil (around each tooth) 115turns
Diameter of each conductor strand 0.63mm (- AWG 22)
Conductor insulating material max. temp. _ 180C
Slot insulating material max. temp. _ 200C (Nomex410 or similar)
Cooling arrangements Natural convection (Fin cooled)
Area per half slot 86.3mm
2
Number of poles 14poles
No. of magnet pieces 28pieces (2 X 14poles)
Magnet dimensions 13.5mm X 2mm X 79mm
Desired remnant flux density (Grade) 1.04 1.09T (Raremag - N24EH)
Magnet temperature rating 200C
Magnet coating Ni-Cu-Ni
Core material Non-oriented silicon steel
Saturation mag. of laminations 1.68T @5000A/m
Predicted core loss at 50Hz/1.5T 2.60W/Kg
Yield strength of laminations _ 300N/mm
2
Resistivity of lamination 50O-cm
Rated voltage (RMS) 240V/phase
Rated current (RMS) 2.55A/phase
Maximum rated torque at base speed 18Nm
Maximum operating speed 6857rpm
d-axis inductance (at rated current) 81.16mH
q-axis inductance (at rated current) 84.45mH
Stator resistance at ambient temperature 10O/phase
Magnet flux linkage 0.48Wb
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
114
4.7.1 Back EMF from the Finite Element Model
The predicted back EMF from the final FE model is shown below:
Fig. 4.24Three-phase induced line to neutral back EMF voltage from the FE model (at 50Hz)
It can be observed that the modelled phase back EMF is near-perfectly sinusoidal, all
three phases are balanced and 120 apart. The induced voltage per phase when the
machine is ran at 428.6RPM (50Hz) is 91.6Vrms. The modelled line to line induced
voltage at the same speed is sinusoidal, and has a value of 158.6Vrms. The linear
relationship between induced line to line voltages versus rotor speed is shown in fig.
4.25:
Fig. 4.25Induced line to line voltage versus speed
-150
-100
-50
0
50
100
150
0 0.005 0.01 0.015 0.02
Va(L-N)
Vb(L-N)
Vc(L-N)
0
200
400
600
800
1000
1200
1400
1600
0 500 1000 1500 2000 2500 3000 3500 4000 4500
I
n
d
u
c
e
d

V
o
l
t
a
g
e

(
V
)
Time (s)
I
n
d
u
c
e
d

V
o
l
t
a
g
e

(
V
r
m
s
)
Speed (RPM)
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
115
To determine the winding factor, the final design was compared to an equivalent
integral-slot, double-layer, DW machine with the same rotor. This comparison is shown
in fig. 4.26 below. The corresponding harmonic spectrums of these waveforms are
shown in fig. 4.27.
Fig. 4.26 Comparison of EMF waveform between the CW-IPM and DW-IPM
Fig. 4.27Comparison of EMF waveform between the CW-IPM and DW-IPM frequency
spectrum
It can be seen that both these winding types produce equally sinusoidal back EMF
waveforms. However, due to a higher winding factor, the DW-IPM machine produced a
higher fundamental term, (233.2Vrms), compared to the CW-IPM machine,
-250
-200
-150
-100
-50
0
50
100
150
200
250
0 0.005 0.01 0.015 0.02
CW-IPM DW-IPM
0
50
100
150
200
250
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
CW-IPM DW-IPM
4 5 6 11 12 13
I
n
d
u
c
e
d

V
o
l
t
a
g
e

(
V
)
Time (s)
I
n
d
u
c
e
d

V
o
l
t
a
g
e

(
V
)

Harmonic number
224.3V
rms
233.2V
rms
Near-zero harmonics
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
116
(224.3Vrms). The DW-IPM machine has a distribution factor of 0.9659, and a pitch
factor of 0.9695; this results in a winding factor of 0.933. Based on these values,
fundamental magnitude achieved for a unity winding factor should be 249.9Vrms and
the expected magnitude of the 18-slot, 14-pole model should be (249.9Vrms x 0.902 =
225.45Vrms). With a value of 224.3Vrms achieved, theresult agrees with the tabulated
values in table 3.1.
4.7.2 Cogging Torque from the Finite Element Model
The cogging torque magnitude of the final design, (3/7Spp), compared to an integral-
slot (2Spp) DW model, is shown in fig. 4.28. It is shown that the final design produces
2.5 times lower peak to peak cogging torque magnitude compared to the DW model.
This is due to the elimination of periodicity of slots and poles, which lowers the
magnitude of the cogging torque but increases its fluctuating frequency.
Fig. 4.28Cogging torque of final design comapred to an equivaent integral-slot DW model
4.7.3 Inductance and Saliency Ratio from the Finite Element Model
Since, the calculation of inductances by aligning the winding axes to the d- and q-axis
of the rotor is not accurate in CW machines, d- and q-axis inductances are calculated by
applying AC standstill test conditions to the FE model. The self- and mutual inductance
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
117
waveforms of the final design from the FE model are shown in Fig. 4.29 (for single-
phase, 50Hz, 3A
rms
current excitation).
Fig. 4.29Self- and mutual-inductance waveform of final model with 3A
rms
current excitation
Ld, Lq and saliency ratio obtained from the DC and second harmonic terms of these
waveforms are as follows.
I
d
= (I
u0
-H
ub0
) -_
I
u2
2
+H
ub2
]
I
d
= 81.16mH
I
q
= (I
u0
-H
ub0
) +_
I
u2
2
+H
ub2
]
I
q
= 84.45mH
=
L
q
L
d
= 1.04
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
118
4.7.4 Torque Performance from the Finite Element Model
In order to account for small fluctuations, steps/electrical cycle had to be small (200
steps per cycle). The characteristic current of the machine was found to be 2.55A
rms
(at
which the widest CPSR could be achieved). With the peak voltage set to 240V
(1-l)
, the
rated torque at a base speed of 428.6rpm is 15.14Nm (equivalent to 679W of power).
The torque ripple at this speed is 2% of the torque produced as shown in fig. 4.30.
Fig. 4.30Torque ripple of final model at base speed
With the abovementioned rated values, a >10:1 CPSR can be achieved. The peak
power achieved was 880W (at 300Hz).
Fig. 4.31CPSR of final design with base frequency of 50Hz
The power versus speed characteristic shown in fig. 4.31 is for the ideal case where
losses are ignored. Due to frequency-related losses, the output power and CPSR may be
affected. The performance with the inclusion of losses will be discussed in chapter 5.
14.7
14.8
14.9
15
15.1
15.2
15.3
15.4
15.5
0 0.01 0.02 0.03 0.04
0
200
400
600
800
1000
0 500 1000 1500 2000 2500 3000 3500 4000 4500
P
o
w
e
r

(
W
)
Speed (RPM)
T
o
r
q
u
e

(
N
m
)

Time (s)
>10:1 CPSR
0.31Nm
(p-p)
Chapter 4: Design of an IPM Machine with Concentrated Windings for Field Weakening Applications
119
4.8 CONCLUSION
This chapter showed the optimisation process of the 14-pole, 18-slot, double-layer CW-
IPM machine to achieve a wide CPSR and desired torque performance. The key steps in
this design were:
Positioning of phase windings to achieve highest winding factor
Material considerations
Rotor magnet design and structural considerations
Optimising the CPSR with by satisfying characteristic current equilibrium
conditions
Further extend the speed range by airgap length and rotor geometry variations
FE analysis showed that thefinal design can achieve over 10:1 (lossless) CPSR with the
capability of producing over 880W of peak power for a voltage limit of 240V
(l-l)
. By
comparing FE models of the final CW-IPM machine with the integral-slot DW-IPM
machine, it was shown that the expected winding factor (0.902) and significantly lower
cogging torque can be achieved. It was also shown that high inductance values were
achieved with L
d
being similar to L
q
resulting in a saliency ratio is almost unity (1.04).
In this chapter, the design process only considered the ideal scenario where both
electrical and mechanical losses were omitted. Losses will be considered separately in
the chapter 5. The test results from the constructed design will be shown in chapter 7,
where the FE results will be verified.
120
CHAPTER 5
ANALYSIS OF LOSSES IN AN IPM MACHINE WITH
CONCENTRATED WINDINGS FOR FIELD WEAKENING
APPLICATIONS
5.1 INTRODUCTION
A key aim in almost all high performance PM machine design is to minimise losses. In
machines used for field weakening applications, frequency related losses in particular
have to be carefully considered.
Losses in PM machines are separated into two main areas electrical losses and
mechanical losses.
Electrical Losses:
Core losses Eddy current and hysteresis losses
Permanent magnet losses
Copper loss
Mechanical losses:
Bearing losses
Windage losses
Chapter 5 will highlight the causes of increased frequency-related losses in an IPM
machine with CW in comparison to DW. The effects of varying material type and
geometry will be investigated. The abovementioned electrical and mechanical losses
will be studied and quantified in terms of the final FE design presented in chapter 4.
Losses would be found at various frequencies and used to provide a more realistic field
weakening performance of the machine.
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
121
5.2 MMF HARMONICS AND LOSSES IN MACHINES WITH
CONCENTRATED WINDINGS
Compared to DW, the MMF waveform produced by CW contains additional MMF
harmonics and sub-harmonics not rotating in sync with the synchronous frequency. Fig.
5.1 show the magnetic flux distribution in a CW and DW machine. (The flux
distribution shown is by armature excitation only).
(a) Flux distribution of 14-IPM rotor with a
fractional-slot CW stator
(b) Flux distribution of 14-IPM rotor with an
integral-slot DW stator
Fig. 5.1 Comparison of flux distribution in a CW and DW machine
MMF (J) produced by the stator coils is basically expressed as flux density produced by
the stator (B
s
) linked across an airgap of surface area (A
rot
) with reluctance (J
g
).
J = B
s
A
ot
J
g

(5.1)
If the machine geometry, including airgap length and slot opening width, is kept
constant, the only parameter that will affect MMF across the airgap is the flux density
produced by the armature reaction. The airgap flux waveforms generated by the DW
model (fig. 5.2), has afundamental component at 7 times the base frequency, which is
the torque producing term in this 14-pole machine model. On the other hand, CW
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
122
produces a dominant fundamental in addition to several other significant harmonic and
sub-harmonic components. Fig. 5.3, gives a comparison of MMF waveforms produced
by single-layer and double-layer CW.
Fig. 5.2 Airgap flux waveform and harmonic spectrum producedfrom aDW model
(a)
(b)
Fig. 5.3 Airgap flux generated by 18-slot (a) single-layer CW, (b) double-layer CW
-0.25
0
0.25
0 100 200
Tesla
0
0.1
0.2
0.3
0 25 50
Tesla
-1
-0.5
0
0.5
1
0 100 200
Tesla
0
0.1
0.2
0.3
0 25 50
Tesla
-0.5
-0.25
0
0.25
0.5
0 100 200
Tesla
0
0.1
0.2
0.3
0 25 50
Tesla
Fundamental
Term
Fundamental
Term
Fundamental
Term
n
th
order leakage
harmonic terms
n
th
order leakage
harmonic terms
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
123
The total loss in the core depends on two main terms hysteresis loss and eddy current
loss as shown in (5.2).
P
coc
= K
c

2
B
mux
2
+K
h
B
mux
q
st

(5.2)
where,
K
c
=Eddy current loss constant
K
h
=Hysteresis loss constant
B
mux
=Maximum flux density
p
st
= Material dependent Steinmetz loss constant (in the range of 1.6 to 2.0)
While hysteresis loss depend largely on material and flux density, eddy current loss is
proportional to the frequency squared, making it much more susceptible to leakage
harmonic components which are several times higher than the operating frequency
[208]. In the field-weakening region, an additional CPSR factor is further multiplied to
the fundamental and leakage components. For example: in operation at 10 times the
base frequency of 50Hz, a leakage harmonic component n=42 would be fluctuating at
3000Hz.
To better visualise the proportion of losses, it is common to divide the total loss by
frequency to keep hysteresis term constant; with the p
st
assigned as 2, (5.2) can be
simplified [192, 197, 209].
P
coc
B
mux
2
= K
c
+K
h
(5.3)
A graphical representation of this equation is shown in fig. 5.4.
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
124
Fig. 5.4 Contribution of eddy current and hysteresis loss
As the hysteresis loss is largely material dependent and little can be done to minimise it,
emphasis will be placed on eddy current loss, in which the choice of lamination
thickness can help in minimising losses.
For a general concept on how the increase in airgap harmonics affects eddy current loss,
Faraday and Maxwell equations can be used.
Equating the electric field (E

) in a closed path c along the surface to the induced voltage


(I
nd
) and flux linkage we get:
_ E

Jl
c
= I
nd
=
J
Jt
_ B

Js =
J
Jt
s

(5.4)
Equating current density ([) to the electric field, we have:
[ = oE (5.5)
Say a fluctuating magnetic flux density (B) acts on a thin piece of material with
conductivity (o), thickness (t) and width (w). Eddy current (P
eddy
) induced in this
conductor can be expressed as follows:
esisloss
Frequency (Hz)
P
coc
B
mux
2

K
h

K
c

t andh
f
1
Contribution of
eddy current
loss
Contribution of
hysteresis loss
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
125
P
cdd
=
1
w
_
[
2
o
JI
(5.6)
Assuming that flux density is uniform across the surface, flux linkage will be:
= Bw (5.7)
The induced voltage in (5.4) can be expressed as:
I
nd
=
J
Jt
= w
JB
Jt

(5.8)
If is reduced to infinitely thin sections (x), the electric field along the closed path in
(5.4) can be described by:
E =
I
nd
2w
= x
JB
Jt

(5.9)
From (5.5) the induced current density at the surface of the material is given by:
[ = oE = ox
JB
Jt

(5.10)
The eddy current loss shown in (5.6) in terms of the induced current density can be
described as follows:
P
cdd
=
1
w
_
1
o
_ox
JB
Jt
]
2
:
2
,
-:
2
,
wJx (5.11)
P
cdd
=

2
o
12
_
JB
Jt
]
2

(5.12)
For integral-slot DW, a sinusoidal airgap flux density (fig. 5.2) is assumed, then (5.12)
can be solved to obtain:
P
cdd
=

2
o
12
_
J
Jt
B
mux
sint]
2

(5.13)
P
cdd
= _

2
o
6
(1 +cos2t)_
2
B
mux
2
(5.14)
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
126
where the terms in the square brackets are usually absorbed in the eddy current constant
(K
e
).
In CW, the function would not only depend on the fundamental but other leakage
harmonic components as well. The eddy current loss for CW can thus be generally
expressed in (5.15) where B
n
is the n
th
harmonic component of the waveform as shown
fig. 5.3.
P
cdd
=

2
o
12
_
J
Jt
B
n

1
n
sin(nt)
N
n=1
_
2

(5.15)
From (5.15) it can be seen that the most effective way to reduce eddy current loss is by
minimising flux density harmonics. Although the introduction of additional harmonic
components is unavoidable with CW, double-layer CW windings substantially reduce
these components. For this particular reason, double-layer was chosen over single-layer
CW windingsin the final design.
In the PM machine, eddy current loss occurs mainly in the core and magnets. As
mentioned in the chapter 4, non-oriented silicon steel is chosen as the core material due
to its price and performance characteristics. Within the range of silicon steel grades,
conductivity may vary up to 35% [198, 210], resulting in different core losses. The
thickness of the material also affects core loss. This chapter will study the effects of
core loss as the steel grade/thickness is varied.
For the magnet material, there is a huge variation of conductivity between the two
different methods by which NdFeB magnets are made. Sintered magnets have a typical
conductivity of 625X10
3
(O-m)
-1
whereas bonded magnets have a conductivity of
7.14X10
3
(O-m)
-1
. Thus, in areas such as in an SPM machine used for FW applications
where magnet losses are high, bonded magnets are commonly used at the expense of
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
127
lower magnet remanent flux densities. In cases where high torque density and low
magnet losses are required, sintered-segmented magnets can be used [140, 211].
The comparison of SPM and IPM magnet losses, as well as the effects of magnet
segmentation and variation of magnet type will be quantified this chapter.
5.3 CORE LOSS
In machines used for field-weakening applications, frequency-related losses have to be
carefully considered and minimised due to constant operation at high frequencies.
Furthermore the increase in leakage harmonic terms as a result of applying CW makes
the machine more susceptible to increased core and rotor losses.
One method of separating the hysteresis and eddy current losses is derived from the
relationship mentioned by Yeadon [212]; the book states that in typical steel
laminations (grades M19 through to M45), hysteresis loss at 60Hz make up
approximately 67% of the total core loss and eddy current loss makes up the other 33%.
The material chosen (35RM300) has core loss 3.25W/kg at 60Hz/1.5T (shown in fig
5.5), which is similar to grade M19. Thus, this approximation is valid for the breakdown
of hysteresis and eddy current loss at 60Hz and the total core loss can be expressed as:
P
coc
= P
cdd
+P
hs
(5.16)
where,
P
cdd
=
1
3
P
coc

(5.17)
P
hs
=
2
3
P
coc

(5.18)
This estimation gives us an eddy current loss of 1.08W/kg and hysteresis loss of
2.17W/kg. In order to determine core losses due to harmonics components, various
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
128
operating frequencies as well as saturation levels, the hysteresis (K
h
) and eddy current
loss constants (K
c
) must first be found.
Fig. 5.5 Core loss curve of 35RM300 provided by Sankey showing core loss at 60Hz/1.5T
To ensure consistency between FE models in Magsoft-Flux2D, a narrow annular-core
specimen consisting of thin laminations is created (as shown in fig. 5.6):
Fig. 5.6 Annular steel model to determine eddy current and hysteresis loss constants
The surface area of the core was chosen such that the flux distribution throughout the
core is at a relatively constant flux density in this case 1.5T. Frequency of excitation
Constant flux density
throughout surface area
(3.3% difference margin)
I
s
(60Hz)
Annular
laminated-core
structure
Coils with number of
turns producing 1.5T
flux density in core
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
129
was set to be 60Hz. The volume of the core was chosen to be equivalent to 100g of
silicon steel.
From this model, the values of K
c
and K
h
was found by trial and error, to achieve eddy
current loss of 1.08W/kg and hysteresis loss of 2.17W/kg. The values were 0.29 and
151 respectively. These loss constants can then be used to obtain the core loss in the
machine model at various excitation frequencies. (K
c
and K
h
is found in this way for
each steel grade).
5.3.1 Comparison of Steel Grades
Here the chosen steel grade (35RM300) will be compared with two other steel grades
with different thicknesses- 50J N350 and 65J N800 [198, 213]. Key features of the
specific grades are shown in table 5.1.
Table 5.1
Properties of compared core grades
Material
Type
Lamination
Thickness
Saturation
Mag. (B
50
)
Stacking
Factor
Total Core Loss
@60Hz/1.5T
K
h
K
c
35RM300 0.35mm 1.68T 95% 3.25W/kg 151 0.29
50J NE350 0.50mm 1.68T 96% 4.45W/kg 263 0.39
65J NE800 0.65mm 1.72T 97% 10.15W/kg 495 0.89
With K
c
and K
h
found using the annular model, the losses at 50 and 500Hz are shown in
fig. 5.7. Stacking factor in the above table was stated as a reference(being proportional
to output torque) it should be noted that in the loss calculations, stacking factor was
not included as differenceswould not be very significant.
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
130
Fig. 5.7 Core loss comparison at 50 and 500Hz with different steel grades
As a percentage of output power, core loss in the constant torque region (50Hz) is low
(<2.5%), therefore a maximum difference in total core losses among the steel grades (of
10.6W) does not significantly affect the efficiency. However, in the higher field-
weakening region (500Hz), the core loss as a percentage of output power becomes
significant (>13%), and the maximum difference in total core losses (of 74.1W) would
result in almost a 10% difference in efficiency.
The results emphasise the importance of using thin laminations with low losses when
designing machines for field-weakening applications. Thicker laminations, (like the
65J N800), are more suitable for machines operating at lower frequencies, (within the
constant torque region), where torque density is of higher priority. As a percentage of
total loss, it is expected that core loss would be low. Hence this rough estimation
indicated above would suffice for the study of this preliminary prototype. A more
detailed measurement of losses would be required for future prototypes where core
losses significantly impact the overall efficiency of the machine.
1.4
4.7
6.0
14.3
19.1
33.4
2.4
8.0
10.3
24.2
29.5
53.7
4.5
15.3
19.7
46.6
60.9
107.5
0
20
40
60
80
100
120
Rotor core
loss @50Hz
Stator core
loss @50Hz
Total core
loss @50Hz
Rotor core
loss @500Hz
Stator core
loss @500Hz
Total core
loss @500Hz
35RM300 50JN350 65JN800
P
o
w
e
r

l
o
s
s

(
W
)

Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
131
5.3.2 Core Loss of Final Design
The core loss in the final design with the chosen material (35RM300) at various
frequencies is shown in fig. 5.8.
Fig. 5.8 Core loss with chosen steel grade at various frequencies
As expected, core loss increases exponentially with frequency. The percentage of stator
to rotor core loss is several times higher (77% from the stator, compared to 23% from
the rotor, at 50Hz) in the constant torque region where the d-axis current is zero. In the
field-weakening region, rotor core losses become substantial, contributing to a large part
(43% at 500Hz) of the total core loss in the machine.
0
5
10
15
20
25
30
35
40
45
50
50 100 200 300 400 500
Rotor Core
Loss
Stator Core
Loss
P
loss(rotor)
=2.57W
P
loss(stator)
=3.73W
P
loss(rotor)
=2.23W
P
loss(stator)
=5.94W
P
loss(rotor)
=8.09W
P
loss(stator)
=9.49W
P
loss(rotor)
=11.12W
P
loss(stator)
=13.89W
P
loss(rotor)
=14.32W
P
loss(stator)
=19.09W
P
o
w
e
r

l
o
s
s

(
W
)

Frequency (Hz)
P
loss(rotor)
=1.36W
P
loss(stator)
=4.68W
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
132
5.4 MAGNET LOSS
With the core loss substantially reduced by the choice of thin silicon steel laminations,
thetimevarying fields may still createsubstantial losses in the magnets, especially in
SPM machines. A great deal of research interest is focused on the study and
minimisation of magnet losses in CW-PM machines. Commonly used strategies to
reduce magnet losses are by the use of bonded instead of sintered magnets at the
expense of lower magnet strength, or the use of sintered-segmented magnets [141, 208,
214, 215].
While the magnet eddy current loss constant can be easily calculated by conductivity
and thickness of the magnets, the hysteresis loss data is not being readily available.
Here, the hysteresis loss constant was estimated by extrapolating measured data from
Fukuma et al. [216] and Shinichi et al. [208]. In their work, hysteresis and eddy current
losses in sintered neodymium were measured when exposed to low flux densities up to
0.1T. The results show that while the hysteresis loss increase was relatively linear; there
was a slight exponential increase to eddy current loss. Extrapolated polynomial trend
lines were fitted into measured results obtained from the abovementioned papers, as
shown in fig. 5.9. More accurate loss measurements will be left for future work when
the equipment becomes available to test magnet losses. For this work, this estimate
would suffice as it will be shown later that magnet losses in the CW-IPM machine make
up only a very small percentage (0.5%) of total losses.
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
133
Fig. 5.9 Extrapolated values of measured hysteresis and eddy current loss for sintered
neodymium magnets at 50Hz [208, 216]
The typical operating point of the PMs in the final design is 0.6T. Values obtained from
the extrapolation at that flux density are 10.1W/kg for hysteresis and 4.6W/kg for eddy
current loss. There are several FE methods for calculating magnet losses. The purpose
of using this method (declaration of eddy and hysteresis loss constants) is due to the
advantage of being able to study the effect of each loss separately.
For neodymium magnets, conductivity of magnets is readily available from
manufacturers datasheets. Thus eddy current loss constant can be calculated by the
following formula [209]:
K
c
= o

2
n
2
6

(5.19)
With each magnet piece being 79mm long, K
c
for sintered and bonded neodymium
magnets are calculated to be 6416 and 73 respectively.
With the relation between K
c
and K
h
obtained from the extrapolated values shown infig.
5.9, K
h
can be estimated. The magnet loss constants can then be used to study magnet
losses with 3D FE model using ANSOFT-Maxwell13.
0
2
4
6
8
10
12
14
0 0.1 0.2 0.3 0.4 0.5 0.6
Hystesis Loss [W/kg] Eddy Current loss [W/kg]
Poly. (Hystesis Loss [W/kg]) Poly. (Eddy Current loss [W/kg])
10.1W
4.6W
P
o
w
e
r

l
o
s
s

(
W
/
k
g
)

Flux Density (T)
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
134
5.4.1 Comparison of SPM and IPM Magnet Losses
Compared to the IPM machine, (shown in fig 5.10), the SPM machine has permanent
magnets that are directly exposed to airgap flux, (shown in fig, 5.11), thus higher
magnet losses are expected [217]. Here, PM losses in SPM machines are quantified and
compared against the IPM machine.
Fig. 5.10 3D model of a single-pole and single-phase excitation v-shaped IPM
Fig. 5.11 3D model of a single-pole and single-phase excitation SPM
In these two models, airgap between the magnet outer surface and stator inner radius is
1.2mm, (the same airgap as in the IPM model). The volume and grade of magnet are
kept constant. A magnet loss comparison between these two machine types at 50Hz and
500Hz are tabulated in table 5.2 and graphed in fig. 5.13 below. The magnet losses
quantified in this section refers to total magnet losses in a 14-polemachine.
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
135
Table 5.2
Comparison of magnet losses between IPM and SPM machine
50Hz 500HZ
P
eddy
P
hys
P
mag(total)
P
eddy
P
hys
P
mag(total)
IPM
(sintered mag.)
0.018W 0.034W 0.052W 0.560W 0.340W 0.900W
SPM
(sintered mag.)
2.133W 4.140W 6.273W 67.320W 37.900W 105.220W
SPM
(bonded mag.)
0.024W 1.110W 1.134W 0.771W 11.094W 11.865W
Fig. 5.13a Magnet losses in an IPM machine with sintered magnets
Fig. 5.13b Magnet losses in an SPM machine with sintered magnets
Fig. 5.13c Magnet losses in an SPM machine with bonded magnets
0.02
0.03
0.05
0.56
0.34
0.90
0
0.2
0.4
0.6
0.8
1
Eddy current
loss @50Hz
Hysteresis
loss @50Hz
Total magnet
loss @50Hz
Eddy current
loss @500Hz
Hysteresis
loss @500Hz
Total magnet
loss @500Hz
P
o
w
e
r

l
o
s
s

(
W
)
2.1
4.1
6.3
67.3
37.9
105.2
0
20
40
60
80
100
120
Eddy current
loss @50Hz
Hysteresis
loss @50Hz
Total magnet
loss @50Hz
Eddy current
loss @500Hz
Hysteresis
loss @500Hz
Total magnet
loss @500Hz
P
o
w
e
r

l
o
s
s

(
W
)
0.0
1.1 1.1
0.8
11.1
11.9
0
2
4
6
8
10
12
14
Eddy current
loss @50Hz
Hysteresis loss
@50Hz
Total magnet
loss @50Hz
Eddy current
loss @500Hz
Hysteresis loss
@500Hz
Total magnet
loss @500Hz
P
o
w
e
r

l
o
s
s

(
W
)
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
136
Because the magnets are being exposed to the airgap flux containing high leakage
harmonic content, (and with no protecting sleeves around the magnets), total magnet
losses in the SPM models are very much higher than that of the IPM model. In [216,
208], the measurement of magnet losses in an SPM machine show that at frequencies
lower than 1kHz, hysteresis loss is higher compared to eddy current loss. Results in
table 5.2 is consistent with the measured losses, in the sense that hysteresis loss is
higher compared to Eddy current loss at 50Hz but lower at 500Hz.
It should also be mentioned that the inverse is true for rotor core losses (that is: the rotor
core loss in the SPM machine is negligible as compared to the IPM rotor core loss due
to the magnets damping most of the varying harmonic terms from armature reaction).
Fig. 5.13 also shows that the total magnet losses with SPM bonded magnets are about
10 times lower than that of the SPM model with sintered magnets.
While the use of bonded magnets is highly favourable in SPM machines, in IPM
machines the use of bonded magnets at the expense of a reduction in torque density is
clearly not, since sintered PM losses make up less than 0.1% of the power produced.
5.4.1 Effects of Magnet Segmentation
Despite the very low magnet loss obtained in the IPM model, it would be useful to study
the effects of magnet segmentation in the SPM and IPM models. The work done in this
section takes reference to well-established work in various literature, some of whichare
[140, 214, 215, 218, 219].
In work done by Yamazaki et al. in [140], magnet losses due to carrier and slot
harmonics in different rotor types were compared. With the variation of the number of
segments, SPM machines show a gradual decrease in manget eddy current loss as the
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
137
number of segments is increased, whereas IPM machines exhibit peaks (shown in fig.
5.14).
Fig. 5.14 Variation of magnet eddy current loss with number of magnet segments due to slot
and carrier harmonics Comparison between IPM, inset and SPM rotors [140]
In this section, the effects of segmenting magnets in the IPM and SPM models are
compared. The same models shown in fig. 5.10 and 5.12 are used. Fig. 5.15 shows the
circulating eddy currents in the single SPM magnet pole piece, as well as in the IPM
magnet pole consisting of two magnet pole pieces.
(a) SPM magnet pole (b) V-shaped IPM magnet pole
Fig. 5.15 Circulating eddy currents in a non-segmented poles
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
138
Fig. 5.16 shows the circulating eddy currents same magnet poles when the magnets are
segmented.
(a) SPM magnet pole (b) V-shaped IPM magnet pole
Fig. 5.16 Circulating eddy currents in a segmented magnet poles
The total magnet losses in all 14-poles obtained at 50 and 500 Hz for the SPM and IPM
machine models are shown in fig. 5.17 and fig. 5.18 respectively.
Fig. 5.17 Magnet losses in an SPM machine with variation of segment number
0
20
40
60
80
100
120
No Segments 3 Segments 6 Segments 9 Segments 12 Segments
SPM @50Hz SPM @ 500Hz
T
o
t
a
l

m
a
g
n
e
t

l
o
s
s

(
W
)

100%
44.4%
41.8%
39.7% 39.6%
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
139
Fig. 5.18 Magnet losses in an IPM machine with variation of segment number
Fig. 5.17 and 5.18 show that, while there is a gradual decrease in PM losses for the SPM
model, the decrease in PM lossesfor the IPM model was not constant, exhibiting a peak
with 9 segments. These results comply with the results obtained by Yamazaki shown in
fig. 5.14. It can be seen that segmentation has greater effect on losses in the SPM
machine as compared to the IPM machine, where more segments are required to achieve
the same percentage of losses.
With a difference of a mere 49.3% (at 500Hz) with 12 magnet segments compared to a
single pole-piece, the cost of increased complexity and duration of manufacturing the
IPM rotor is not worthwhile for this design.
5.4.2 Magnet Losses in the Final Design
In the final sintered non-segmented magnets were used. The total magnet losses for the
final model us shown in fig. 5.19:
0
0.2
0.4
0.6
0.8
1
No Segments 3 Segments 6 Segments 9 Segments 12 Segments
IPM @50Hz IPM @ 500Hz
T
o
t
a
l

m
a
g
n
e
t

l
o
s
s

(
W
)

100%
70.9%
59.2%
78.8%
50.7%
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
140
Fig. 5.19 Magnet losses versus frequency in the final design with sintered NdFeB magnets
The above figure shows that the total magnet loss for the final design is very low
(making of for less than 1.5% of total output power of the machine). The breakdown of
losses show that hysteresis loss increases linearly, and is the dominant loss in the
magnet for frequencies <200Hz, whereas the eddy current loss increases exponentially
with frequency due to its frequency squared dependence, and begins to dominate at
higher frequencies. Compared to total core lossin the rotor, magnet losses in the IPM is
very small. Majority of the losses occurs in the rotor iron.
5.5 STATOR WINDING LOSS
Stator winding loss also known as I
2
R, copper or joule loss occurs when the
armature windings are excited by an external source. Of the total loss in PM machines,
the largest portion is usually due to I
2
R loss [220]. I
2
R is not frequency dependent and is
constant throughout the speed range, so the CPSR is not affected by this loss. This is
due to the copper conductors being thin enough to have 100% skin-depth throughout the
operating region. I
2
R lossis described in the following formula:
I
2
R = 2N
coI
I
2
p
l
c]]
A
w
(5.20)
where,
N
coI
=Number of turns per coil
p =Conductor resistivity (1.68X10-8 for copper)
A
w
=Cross-sectional area of wire
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
0 50 100 200 300 400 500 600
Eddy current
loss
Hysteresis
loss P
o
w
e
r

l
o
s
s

(
W
)

00 300 40
Frequency (Hz)
P
eddy
=34%
P
hys
=66%
P
eddy
=51%
P
hys
=49%
P
eddy
=59%
P
hys
=41%
P
eddy
=64%
P
hys
=36%
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
141
A common standard for wire sizes is the American wire gauge (AWG) standard. Some
wire sizes for typical machines in the range of 1kW are shown in table 5.3 below [221,
222]. Values indicated are for a 25C ambient temperature and for frequencies below
60Hz.
Table 5.3
Suitableconductor sizes properties
AWG
Diameter
(mm)
Area
(mm
2
)
Copper Resistance
(O/km)
17 1.150 1.04 16.61
18 1.024 0.823 20.95
19 0.912 0.653 26.42
20 0.812 0.518 33.31
21 0.723 0.410 42.00
22 0.644 0.326 52.96
23 0.573 0.258 66.79
24 0.511 0.205 84.22
25 0.455 0.162 106.2
26 0.405 0.129 133.9
In machines withhigh torque densities, I
2
R losses are greater due to the larger amount
of current and number of turns (ampere-turns AT) required. One way of reducing this
loss is by the use of thicker conductors, which would in turn require a larger slot area.
The slot area in the stator is maximised by optimising tooth and yoke area, such that the
flux densities in these areas are kept just under the saturating limit of the core material,
as shown in chapter 4.
With a fixed slot area, the only way to increase conductor size and lower I
2
R losses is
by increasing the slot-fill factor S
]]
given by the following equation:
S
]]
=
N
coI
A
w
A
s
(5.21)
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
142
In the final CW-IPM design, a slot-fill factor of 41.5% was assumed.
The estimated average winding length is shown in fig. 5.20. The average length of each
end potion of the winding is 10mm. By adding these to the 80mm stator stack length,
the estimated axial length of each strand is 100mm. The span of the slot measured from
the inner diameter of the stator yoke is approximately 21mm. Therefore the total
estimated length per turn is 242mm.
Fig. 5.20 Inner-diameter view of stator teeth showing estimated winding length
Each coil consists of 115turns, and each phase consists of 6 coils. Therefore the total
number of turns per phase is 690. The total length of wire required for 690 turns is
167m. With a 41.5% slot-fill factor, a suitable wire size would AWG22, (which is a
resistance oI 52.96O/km obtained Irom table 5.3). The total calculated resistance for
167m of wire with size is 8.8O/phase at a temperature oI 25C and less than 60Hz
operating frequency.
100mm
Estimated
winding axial
length
80mm
Stator
stack
length
21mm
Estimated slot span
(measured from yoke)
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
143
5.6 MECHANICAL LOSSES
Mechanical losses consist of mainly bearing and windage losses [197, 220, 223].
Bearing loss is dependent on factors such as the bearing type, bearing diameter, rotor
speed, load and lubricant used. Windage loss occurs when friction is created with the
rotating parts of the machine and the surrounding air.
Bearing losses can be calculated with the following formula [220]:
P
bcung
= 0.5
m
k
b
F
b
(5.22)
where,

m
=Mechanical speed of the rotor
k
b
=Bearing loss constant
F =Force acting on the bearing

b
=Bearing inner diameter
The following figure shows the final drawing of the entire rotor. Its mass is calculated
based on the known volume and density of the various materials used. Bearing friction
loss occurs in two bearings.
Fig. 5.21 Drawing of rotor indicating key components to calculate bearing friction loss
Bearing 1
Inner Diameter =20mm
Bearing constant =0.0015
Bearing 2
Inner Diameter =17mm
Bearing constant =0.0015
Rotor core
Density=7650kg/m
3
Total Mass =2.182kg
Magnets
Density=7500kg/m
3
Total Mass =0.448kg
End plates
Density=8000kg/m
3
Total Mass =0.30kg
Shaft
Density=8000kg/m
3
Total Mass =0.650kg
Downward force
acting on bearing
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
144
From the estimated weight of the rotor, the downward force acting on the bearings is
calculated by:
F = mg (5.23)
where,
m =mass of the rotor
g =gravity (9.81m/s
2
)
The power loss due to friction in each bearing is shown in table 5.4 and the total loss is
plotted in fig. 5.22. The calculated forceF and bearing loss constant k
b
are 35.12Nm
and 0.0015 respectively.
Table 5.4
Bearing friction loss on each bearing at various speeds
Frequency 50 100 200 300 400 500 600 700
Speed (rpm) 429 857 1714 2571 3429 4286 5143 6000
P
(bearing1)
(W) 0.020 0.040 0.080 0.121 0.161 0.201 0.241 0.281
P
(bearing2)
(W) 0.024 0.047 0.095 0.142 0.189 0.236 0.284 0.331
P
(total)
(W) 0.044 0.087 0.175 0.262 0.350 0.437 0.525 0.612
Fig. 5.22 Power loss due to bearing friction at various speeds
The results show that bearing loss increases linearly with speed. This loss is small
compared to other electrical losses.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0 1000 2000 3000 4000 5000 6000
Speed Vs
Bearing loss
T
o
t
a
l
B
e
a
r
i
n
g

l
o
s
s

(
W
)

Speed (rpm)
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
145
Windage loss (P
wndugc
) is another form of mechanical loss; however, unlike frictional
loss, it has a non-linear increase in magnitude as speed increases. The rotor can be
modelled as a rotating cylinder in an enclosure, and power loss is due to the resisting
drag torque on the cylinder. This can be expressed as follows [220]:
P
wndugc
= 0.03125
m
3
nk
ct
k

p
u

4
l

(5.24)
where,
k
ct
=Torque coefficient which has to be separately determined
k

=Roughness coefficient (Between 1-1.4)


p
u
=Density of air (1.184 kg/m
3
)

=Rotor diameter
l

=Rotor length
To determine the torque coefficient (k
ct
), the Couette Reynolds number first has to be
determined. The Reynolds number (N
Rc
) is given by:
N
Rc
=
p
oir

l
ug
2p
s(u)
(5.25)
where,
l
ug
=Airgap length
p
s(u)
=Dynamic viscosity of air (18.6 Pa)
and for N
Rc
between 64 and 500, the torque coefficient is given by:
k
ct
= 2
(2l
ug
/

)
0.3
(N
Rc
)
0.6
(5.26)
And for N
Rc
between 500 and 50000, the torque coefficient is given by:
k
ct
= 1.03
(2l
ug
/

)
0.3
(N
Rc
)
0.5
(5.27)
These calculated values are shown in table 5.5, and power loss due to windage versus
speed is plotted in fig. 5.23.
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
146
Table 5.5
Reynolds number, torque coefficient and winding loss at various speeds
Frequency 50 100 200 300 400 500 600 700
Speed (rpm) 429 857 1714 2571 3429 4286 5143 6000
N
Rc
137 274 548 822 1096 1371 1645 1919
k
ct
0.036 0.024 0.015 0.013 0.011 0.010 0.009 0.008
P
windage
(W) 0.001 0.007 0.037 0.101 0.208 0.363 0.572 0.842
Fig. 5.23 Power loss due to windage at various speeds
Fig. 5.23 shows that although windage loss is almost negligible at low speeds, it
increases exponentially and becomes more significant at speeds in the wider field
weakening region.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
0 1000 2000 3000 4000 5000 6000
Speed Vs
Windage loss
W
i
n
d
a
g
e

l
o
s
s

(
W
)

Speed (rpm)
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
147
5.7 MODELLED FIELD-WEAKENING PERFORMANCE WITH INCLUSION
OF LOSSES
The lossless field-weakening performance of the final CW-IPM machine model shown
in chapter 4 (fig. 4.27), achieved a 10:1 CPSR with peak power of 880W. Fig. 5.24.
shows the abovementioned field-weakening performance, with the inclusion of total
losses highlighted in this chapter.
Fig. 5.24 Modelled power versus speed performance with inclusion of losses
With the inclusion of losses, an efficiency of between 78.9% and 73.8% was achieved.
The low efficiency was due to the structure of themachine (long axial length) for the
purpose of fitting into the same casing as comparison machines. It should be noted that
amachineof the same volumebut with awider outer diameter and a shorter axial length
wouldlead toa high efficiency. This will be shown later in chapter 8.
The 10:1 CPSR was maintained but peak power dropped to 700W.
At base speed 50Hz (429rpm), almost all the losses are I
2
R loss (96.6%) and the
remaining 3.4% is due to core loss. Magnet and mechanical losses at lower frequencies
are negligible. At 500Hz (4285rpm) the percentage of core loss is increased to 16.1%.
Magnet and total mechanical losses were still very small (0.4% altogether). A large
percentage of losses (83%) is still due to I
2
R losses.
0
200
400
600
800
1000
0 500 1000 1500 2000 2500 3000 3500 4000 4500
P
o
w
e
r

(
W
)

Speed (rpm)
Input Power Output Power
CPSR =10:1
q = 73.8%
q = 78.8%
q =78.9% q =78.4%
q = 76.7%
q = 74.8%
Chapter 5: Analysis of Losses in an IPM Machine With Concentrated Windings for Field Weakening
Applications
148
5.8 CONCLUSION
This chapter showed how the increase in harmonics resulting from CW affected
frequency-related losses. The two main categories of losses were studied and quantified
based on known material properties as well as theory and experimental results obtained
from available literature. The losses that were covered were:
Electromagnetic losses
Core loss (varies exponentially with frequency)
Magnet loss (varies exponentially with frequency)
I
2
R Loss (constant)
Mechanical losses
Bearing friction loss (varies linearly with speed)
Windage loss (varies exponentially with speed)
With appropriate selection of steel thickness/grade, frequency related losses in this
model remained low, resulting in total core loss of <4.5% of total power at the
maximum operating speed.
In a comparison of magnet losses, it was shown that the magnet losses in the CW-IPM
machine was almost negligible magnet losses as compared to the magnet losses in the
CW-SPM machine. It was confirmed that for the CW machine, the segmentation of
magnets was more effective in reducing SPM as compared to IPM magnet losses.
The overall losses were combined and applied to the lossless FE model, (shown in
chapter 4). It was shown that a 10:1 CPSR was achieved and efficiencies between
78.9% and 73.8% were obtained throughout the speed range.
In chapter 7, the loss prediction methods in this chapter will be experimentally verified
by the efficiency measured from the constructed CW-IPM prototype.
149
CHAPTER 6
VECTOR CONTROL OF THE IPM MACHINE WITH
CONCENTRATED WINDINGS
6.1 INTRODUCTION
Prior to the change of variables proposed by Park [224] in 1929, complex differential
equations were used to describe a machines performance with sinusoidally varying
quantities currents, voltages, and flux linkages.
Park introduced a method (Parks transformation) to refer the time-varying quantities
from thestator to the rotor reference frame, thus allowing these time-varying quantities
to be expressed in terms of DC quantities the direct-axis (d-axis) and quadrature-axis
(q-axis) as shown in fig 6.1. The d-axis, otherwise known as the pole or flux axis, and
has to be aligned with the magnetic axis of phase A. The q-axis, otherwise known as the
inter-pole axis has a 90 electrical degrees displacement from the d-axis.
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
150
Fig. 6.1 2-pole IPM machine showing d-q axis reference frames
In the d- and q-axis reference frame, the voltage and currents are represented by space
vectors as shown in fig. 6.2. Control of the amplitude and phase of these vectors
determines the performance of the machine, thus the term vector control. As vector
control affects the spatial orientation of the rotor and stator fields, it is also commonly
referred to as field-oriented control (FOC) [174, 225, 226]. FOC is a well-known and
simple algorithm which can be easily implemented on a fixed point digital signal
processor.
Phase A axis (r=u)
(Stator ref. frame) (( Direct axis
(Rotor ref. fame)
Phase C axis
(Stator ref. frame)
Phase B axis
(Stator ref. frame)
Quadrature axis
(Rotor ref. frame)
r

Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
151
Fig. 6.2 Space vector dq-axis phasor diagram
With vector control, current is decoupled into torque- (iq) and flux- (id) producing
components. This results in transient response characteristics similar to those of a
separately excited DC machine [227].
The aim of this work is to run the machine over a wide CPSR, as well as to achieve
error-free speed responses under steady state condition. To do so, the vector control
method is employed. A well-known field-weakening method proposed by Morimoto in
[43] was successfully implemented in available DW-IPM machines, however these
methods resulted in a slight over-suppression of the permanent magnet flux in the
CW-IPM machine.
This chapter will state the vector control method, controller architecture and inversion
technique used to produce the final three-phase inputs to the prototype CW-IPM
machine. The id trajectories during field-weakening operation calculated by
equations proposed by Morimoto will be compared with trajectories obtained by
repetitive testings.
d-axis
(flux axis)
q-axis
(EMF axis)
Iq
Ef
IqR
jIqXs
jIdXs
IdR
V
I
Id
y

(f
m
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
152
6.2 CONTROL METHODOLOGY
6.2.1 Basic Equations describing the PM Machine
The 3-phase voltage equation used for describing a PM machine involves stator currents,
flux linkages and reactance values expressed as follows [227-229]:
I
u
= Ri
u
+
J
u
Jt
= Ri
u
+I
s
Ji
u
Jt
+
J
pm(u)
Jt

(6.1)
I
b
= Ri
b
+
J
b
Jt
= Ri
b
+I
s
Ji
b
Jt
+
J
pm(b)
Jt

(6.2)
I
c
= Ri
c
+
J
c
Jt
= Ri
c
+I
s
Ji
c
Jt
+
J
pm(c)
Jt
(6.3)
where,
I
u,b,c
=Phase voltages
i
u,b,c
=Phase currents

u,b,c
=Total flux linkage in each phase windings
I
s
=Synchronous inductance

pm(u,b,c)
=Flux linkage in each phase windings due to rotor field
R =Stator winding resistance per phase
Parks transformation which transforms the three phase quantities to two axisquantities
can be expressed by matrix A:
A =
2
3
l
l
l
l
l
lcos0 cos _0 -
2n
3
] cos _0 -
4n
3
]
sin0 sin _0 -
2n
3
] sin _0 -
4n
3
]
1
2
1
2
1
2
1
1
1
1
1
1
(6.4)
Applying Parks transformation matrix to the three phase voltage, flux and current gives
us the following voltage equations in the dq reference frame.
I
d
= Ri
d
+I
d
Ji
d
Jt
-
c
I
q
i
q

(6.5)
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
153
I
q
= Ri
q
+I
q
Ji
q
Jt
+
c
I
d
i
d
+
c

pm

(6.6)
I
d
= Ri
0
+
J
0
Jt
= 0
for the case where 3-phases are balanced (6.7)
I
d
and I
q
expressed in matrix form is given by:
_
I
d
I
q
_ = _
R -
c
I
q

c
I
d
R
_ _
i
d
i
q
_ +_
I
d
J
Jt
I
q
J
Jt
_ _
i
d
i
q
_ +_
0

pm
_
(6.8)
where,
I
d,q
=d and q axes phase inductances
i
d,q
=d and q axes phase currents

c
=Speed in electrical radians per second

pm
=Peak permanent magnet flux linkage
For an IPM machine, the well-known torque equation derived in [39, 230], in the dq
reference frame is given by:
I
dc
=
3
2
p|
pm
i
q
+(I
d
-I
q
)i
d
i
q
]
(6.9)
It should be noted that saturation casues variation in I
q
and temperature causes
variation in
pm
. However within the operating range of the CW-IPM machine in this
thesis, the rotor and stator steel are operated below the saturation region; hence these
issues would not be studied here.
The parameters of the modelled CW-IPM machine required for torque calculation are:

pm
=
L
`
0(l-n)
/o
m
p
= 0.41 Wb
I
q
= 84.45 mH
I
d
= 81.16 mH
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
154
Here torque will be split into two terms the alignment term and the saliency term as
shown in fig. 6.3, where the variation of torque versus I
d
will be shown. I
q
will not be
shown but will follow the relationship I
q
=
_
I
s
2
-I
d
2
.
Fig. 6.3 Calculated torque comprising of alignment and reluctancetorque from (6.9)
From fig. 6.3 it can be seen that in the CW-IPM model, the contribution of the
alignment torque makes up most of the total torque produced by the machine. Since the
reluctance term is almost negligible, the torque equation maybe simplified to that of an
SPM machine:
I
dc
=
3
2
p|
pm
i
q
]
(6.10)
6.2.2 Variation of Current Phase Angle
The current phase angle (y) is the angle between the current phasor and the back EMF
axis as shown previously in Fig. 6.2. Variation of this angle affects the values of the d-
and q-axis current, which in turn affects the flux and torque produced by the machine.
For control of the prototype CW-IPM machine, the current angle will be varied to
achieve maximum torque per unit current (MTPC) control, where the current angle is
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
155
in-phase or almost in-phasewith the back EMF phasor (fig. 6.4a), and field-weakening
(FW) control, where the current angle is made to lead the back EMF phasor (fig. 6.4b).
(a) Current phase angle under maximum torque per unit current operation
(b) Current phase angle under field-weakening operation
Fig. 6.4Current phasor under MTPCand Field-weakening operation
In 3-phase quantities, the current angle is simply a time delay added into the current
equations as follows:
I
u
= I
mux
sin(2nt +y) (6.11)
I
b
= I
mux
sin _2nt -
2n
3
+y]
(6.12)
time
V
Ef
I
Ef
IqR
jIqXs
V
I

pm
Ef
V
I
time
y
Iq
Ef
IqR
jIqXs
jIdXs
IdR
V
I
Id
y

pm
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
156
I
c
= I
mux
sin _2nt +
2n
3
+y]
(6.13)
With variation of this current angle, the machine can be controlled to operate with
optimal efficiency, maximised torque and a wider CPSR above base speed under the
constraints of current and voltage limits.
For the CW-IPM machine model, the variations of EMF, current and voltage with time
at base speed, (MTPC operation), as well as maximum speed, (FW operation), is shown
in fig. 6.5a and 6.5b respectively.
(a) Respective waveforms at base speed under MTPCoperation
(b) Respective waveforms at maximum speed under field-weakening operation
Fig. 6.5 Back EMF, induced current and induced voltage waveforms under MTPCand
maximum field-weakening operation
y = 79
I
Ef
V
I
Ef
V
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
157
Fig 6.5a shows the current in phase with the emf waveform. While in fig. 6.5b current
leads. This leading current waveform suppresses the back emf induced in the coils.
This suppression also causes distortion in the voltage waveform as shown in the figure.
6.2.3 Current and Voltage Limits
The maximum torque produced by the machine depends on the imposed armature
current limits, the maximum speed is limited by the maximum output voltage of the
inverter. In order for the CW-IPM machine to achieve a wide constant power speed
range and optimal torque density, the operating limits of the drive should be determined
by the circle diagram [31, 230]. The circle diagram consists of current-limiting circles
and voltage-limiting ellipses as shown in fig 6.6. Fig. 6.6a and 6.6b show limiting
values and trajectories for the IPM machine and SPM machines respectively.
(a) Circle diagram for an IPM machine
(b) Circle diagram for a SPM machine
Fig. 6.6 Circle diagram for IPM and SPM machineshowing current and voltage limits of the
system
Current limiting
circles
MTPC trajectory

i
d
y
i
q
u=ubase
u=u1
u=u2
Voltage limiting
ellipses
MT
Field-weakening
trajectory
_-

pm
L
d
, 0]
Is(max)

Current limiting
circles
C
MTPC trajectory

i
d
i
q
u=ubase
u=u1
u=u2
Voltage
limiting
ellipses
u
Field-weakening
trajectory
_-

pm
L
d
, 0]
Is(max)

Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
158
Since the saliency ratio of the CW-IPM machine is so low, the i
d
and i
q
current
trajectories would be similar to that of the SPM machine in fig 6.5b.
In the circle diagrams, the outer most limit of the current-limiting circle is expressed
simply as:
I
s(mux)
= _i
d
2
+i
q
2

(6.14)
and the voltage-limiting ellipse at various speeds is given by:
I
s(mux)
= _I
d
2
+I
q
2

(6.15)
where under ideal and steady state operations I
d
and I
q
derived earlier in (6.5) and
(6.6) is given by:
I
d
= -
c
I
q
i
q
(6.16)
I
q
=
c
I
d
i
d
+
c

pm

(6.17)
Thus, in terms of current values, the voltage limiting ellipse is given by:
I
s(mux)
=
c
_
(
pm
+I
d
i
d
)
2
+(I
q
i
q
)
2

(6.18)
The centre of the voltage limiting ellipses lies at the point:
_-

pm
I
d
, 0] (6.19)
From standstill to base speed, the machine operates within the current limit along the
MTPC trajectory. When base speed is reached, the operating point follows anti-
clockwise along the current limiting circles. This results in an increase in -i
d
at the
expense of i
q
. An increase in -i
d
creates a larger opposing flux to the rotor pole axis,
which temporarily weakens the field of the magnets. Field-weakening of the magnets
helps to limit the amount of flux that is linked to the stator windings, thus limiting the
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
159
back EMF generated. This helps to maintain voltage under operating limits as speed
continues to increase past base speed.
The control limits on the field-weakening range of the machine depend on the
characteristic current. Practical IPM machines can be classified in two categories:
Type 1: |I
s(max)
| <

pm
L
d

Type 2: |I
s(max)
| >

pm
L
d

In terms of field-weakening control, type 1 machines have a finite field-weakening
range, whereas type 2 have an infinite range. This is due to the centre of the voltage
limiting ellipse lying outside the current limiting circle for type 1 machines and inside
for type 2 as shown in 6.7a and 6.7b respectively. With the centre of the voltage limiting
ellipse lying outside of the current limiting circle, there exists a maximum speed beyond
which both the current limits and voltage limits can no longer be satisfied. Whereas if
the centre of the voltage limiting ellipse lies inside the current limits, both these limits
would always be satisfied with the application of the voltage-limited maximum output
trajectory, thus theoretically resulting in an infinite speed range.
(a) Type 1 machine with centre of voltage limiting ellipse lying outside of the current
limiting circle
Field-weakening trajectory

umax
ubase
MTPC
trajectory
_-

pm
L
d
, 0]
i
d
i
q
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
160
(b) Type 2 machine with centre of voltage limiting ellipse lying inside of the current
limiting circle
Fig 6.7 Classificationof machine type by characteristic current
The CW IPM machine falls under the category of a type 1 machine as the rated current
is lower than the characteristic current. Thus the voltage-limited maximum output
trajectory cannot be applied to this machine. For this work the machine would operate
only with the two trajectories- the MTPC trajectory and the field-weakening trajectory.
The i
d
and i
q
current trajectory plotted for the final CW-IPM FE model is shown in fig.
6.8.
Fig. 6.8 i
d
and i
q
field-weakening current trajectory for the CW-IPM model
umax =
Voltage-limited
trajectory
Field-weakening
trajectory
MTPC
trajectory
ubase
i
d
i
q
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
161
Fig. 6.8 shows the very large requirement of negative d-axis current for initial field-
weakening up to 200Hz. The requirement of additional negative d-axis current
decreases exponentially as speed increases. In the FE model, the limit on this field-
weakening trajectory as speed increases is where the armature field can no longer
maintain constant power, (higher than power produced at base speed), at a specific
voltage limit.
6.2.4 Maximum Torque Per unit Current and Field-Weakening Trajectories
Here two trajectories will be compared: firstly, the trajectory calculated by widely used
Morimotos equations in [43]; secondly, values obtained when the machine is regarded
as an SPM machine during MTPC operation and subsequently by repeated testing in the
field-weakening region.
Under MTPC operation, i
d
calculated by [43] is shown in (6.20), and the remaining i
q
to produce torque at full load is shown in (6.21).
i
d
= -

pm
4(I
q
-I
d
)
-_

pm
2
16(I
q
-I
d
)
2
+
I
s
2
2

(6.20)
i
q(mux)
=
_
i
s
2
-i
d
2

(6.21)
When regarded as an SPM machine, the maximum torque point will intersect with the
current limit circle along the q-axis. Thus i
d
will be equal to zero and at full load, i
q
will
be equal to the supply current.
i
d
= 0
(6.22)
i
q(mux)
= i
s
(6.23)
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
162
For both these trajectories, the speed at the maximum torque point is determined by
i
q(max)
as well as the voltage limit of the inverter I
s(max)
.

busc
=
I
s(mux)
_
(
pm
)
2
+(I
q
i
q(mux)
)
2

(6.24)
In the field-weakening region, cross-coupling effects will be omitted, and i
d
will be
regarded as an independent quantity of i
q
. While i
q
is largely determined by the speed
controller to produce sufficient torque to achieve the desired speed, i
d
will be used to
supress flux produced by the magnets and maintain constant power.
From [43], i
d
for field-weakening is calculated by (6.25).
i
d
= -

pm
I
d
+
1
I
d
_
I
om
2
n
s
2
-(I
q
i
q
)
2

(6.25)
Through repetitive testings, the values of i
d
required to weaken the magnet field are
lower as compared to (6.25). The trajectories comparing these two methods applied to
the CW-IPM FE model are shown in fig. 6.9.
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
163
Fig. 6.9 Comparison of d-axis current trajectories by Morimotos equations and through
repetitive testings
From the fig. 6.9it is shown that the CW-IPM model requires a lower amount of d-axis
current to maintain constant voltage/power, as speed is increased. Chapter 7 will show
that a similar characteristic is seen in theconstructed prototype machine.
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
164
6.3 CONTROLLER ARCHITECTURE
The input to the controller is a speed reference signal (e
ref
), and the outputs of the
controller are the desired three-phase voltages and currents with desired magnitudes and
phase displacements. To obtain the desired magnitudes and phase displacements, a 3-
phase fully-controlled inverter is used. The controller block diagram is shown below:
Fig. 6.10 Vector control system block diagram
This system requires a position sensor, which is mounted to the shaft of the machine.
The position feedback is used to convert 3-phase quantities to dq-axis quantities; it is
also differentiated with respect to time to provide a speed feedback. Simple high-gain PI
controllers are implemented to achieve the desired speed and dq-axis current dynamics,
as well as to remove any tracking or following error. The compensated errors must be
first converted back to three-phase references to be fed into the inverter.
Speed
controller
Current ref.
Generator
iq current
controller
id current
controller
dq
-1
dq
Inverter
J
Jt

uref
u
C
Tref
ag
iq(ref)
id(ref)
iq
id
Vref(a,b,c)
ia,b,c
m
CW IPM
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
165
6.3.1 Three-phase Inversion Technique
For traction applications, the only available energy source is usually the battery
producing DC voltages. This DC supply is then inverted to desired 3-phase supply
currents and voltages with desired magnitudes and phase displacements. Fig. 6.11gives
an example of the inverter architecture for the drive system used in this work. (A more
detailed schematic, as well as its connections to the controller board can be seen in
appendix F).
Fig. 6.11 Rectifier Inverter for producing three-phase outputs to the machine
In the experimental set up, a 3-phase AC source is rectified using an uncontrolled full-
bridge rectifier to produce a desiredDC bus input voltage to the inverter. This DC bus
voltage will then be converted to desired AC signals by modulating techniques. Two
common modulating techniques include the sinusoidal pulse width modulation (SPWM)
scheme, and the space vector modulation (SVM) scheme. Here, SVM is preferred due
Controller
ia,b,c
m
S1
S2
S3
S4
S5
S6
VDC
+
-
Switching signals S1 S6
CW IPM
Uncontrolled full-
bridge rectifier
Three-
phase
input
from
source
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
166
to its lower hardware requirements and a higher modulation ratio of 0.907 compared to
0.785 in the SPWM scheme.
In the SVM scheme, the desired three phase quantities V
an
, V
bn
and V
cn
are sampled in
time at a specified sampling frequency (f
s
) and are represented in terms of space vectors
as shown in fig. 6.12.
There are eight possible switching states (V
0
to V
7
), six of which are termed switching
vectors (V
1
to V
6
) and the other two (V
0
and V
7
) are termed zero vectors.
Fig 6.12 Switching vectors of the space vector modulation method
The SVM method uses a look-up table (table 6.1) to determine the switching states.
This look-up table contains a set of switching rules which enable the voltage vector Vx to
rotate continuously with smooth transition from one sector to the next.
V1 (1,0,0)
V2 (1,1,0) V3 (0,1,0)
V4 (0,1,1)
V5 (0,0,1) V6 (1,0,1)
Sector 1
Sector 2
Sector 3
Sector 4
Sector 5
Sector 6
v1
v2
Se
Vx
u
V0 (0,0,0)
V7 (1,1,1)
Vref(max)
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
167
Table 6.1
Space vector modulation look-up table
State S
1
S
2
S
3
S
4
S
5
S
6
V
an
V
bn
V
cn
Space Vector
0 OFF ON OFF ON OFF ON 0 0 0 V
0
=0,0,0
1 ON OFF OFF ON OFF ON
2v
DC
3
-
v
DC
3
-
v
DC
3
V
1
=1,0,0
2 ON OFF ON OFF OFF ON
v
DC
3
v
DC
3
-
2v
DC
3
V
2
=1,1,0
3 OFF ON ON OFF OFF ON -
v
DC
3
2v
DC
3
-
v
DC
3
V
3
=0,1,0
4 OFF ON ON OFF ON OFF -
2v
DC
3
v
DC
3
v
DC
3
V
4
=0,1,1
5 OFF ON OFF ON ON OFF -
v
DC
3
-
v
DC
3
2v
DC
3
V
5
=0,0,1
6 ON OFF OFF ON ON OFF
v
DC
3
-
2v
DC
3
v
DC
3
V
6
=1,0,1
7 ON OFF ON OFF ON OFF 0 0 0 V
7
=1,1,1
Taking sector 1 as an example, the relationship between V
x
and the two vectors V
1
and
V
2
is given by:
I
x
sin (
n
3
-o) = :
1
sin
n
3

(6.26)
I
x
sin o = :
2
sin
n
3

(6.27)
In terms of v
1
and v
2
,
:
1
=
2
V3
I
x
sin (
n
3
-o)
(6.28)
:
2
=
2
V3
I
x
sin o
(6.29)
Vectors v
1
and v
2
, which are subsidiary vectors of vectors V
1
and V
2
respectively,
determines the position of the rotating vector V
x
. The desired magnitude of the voltage
V
x
is controlled by activating vectors V
1
and V
2
for durations t
1
and t
2
respectively over
half a period T
s/2
. The expression of V
x
in terms of activations times is given as:
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
168
I
x
= :
1
+:
2
= I
1
t
1
I
s/2
+I
2
t
2
I
s/2
+(I
0
or I
7
)
t
0
I
s/2

(6.30)
where,
t
1
=
:
1
I
1
I
s/2

(6.31)
t
2
=
:
2
I
2
I
s/2

(6.32)
t
0
= I
s/2
(1 -t
1
-t
2
) (6.33)
The switching pattern for sector 1 is shown in fig. 6.13:
Fig. 6.13 Switching pattern for sector 1
This process is repeated for the other five sectors. In this way, the magnitude, frequency
and phase of the voltage can be varied according to the desired output voltage signal as
specified by the controller.
V0 V1
V2 V7 V7 V2 V1 V0
t0/2 t1
t2 t0/2
Ts/2
Ts = 1/fs
Phase A
Phase B
Phase C
Chapter 6: Vector Control of the IPM Machine with Concentrated Windings
169
6.4 CONCLUSION
This chapter has illustrated the control methodology of the CW-IPM machine. It
showed the voltage and current limits that the drive is subjected to by use of the circle
diagrams. Due to the centre of the voltage limiting ellipses lying outside of the current
limiting circle, the voltage limited trajectory is not applicable for the CW-IPM model.
The model was subjected to two different i
d
current trajectories: one being the widely-
used trajectory calculated by Morimotos equations and the other through repetitive
testings. It was shown that the actual i
d
current required to weakening the magnet field
and maintain constant voltage is lower than that calculated by available equations.
Lastly this chapter also showed the general controller architecture and the SVM
technique used to generate the three phase input quantities.
170
CHAPTER 7
CONSTRUCTION AND PERFORMANCE ANALYSIS OF THE
CONCENTRATED WINDING IPMMACHINE PROTOTYPE
7.1 INTRODUCTION
The work done in previous chapters provided a study of the CW-IPM machine for use
in field-weakening applications. From this study, a final 14-pole, 18-slot model was
designed and optimised using FE analysis. The final FE model was built to verify
studies and predicted performance characteristics.
This chapter will first give an overview of the construction process of the CW-IPM
prototype (which was the final FE design shown towards the end of chapter 4).
Problems encountered and lessons learnt during the construction process will be stated
in order to facilitate quicker and less problematic manufacturing for future designs.
Subsequently, the open circuit parameters of the prototype machine will be measured,
and compared with results achieved by the FE model. Control techniques shown in
chapter 6 will be used to run the machine in constant torque and field-weakening
regions. A steady state analysis will be performed and the torque performance of the
CW-IPM machine in both operating regions will be shown. Transient characteristics of
the machine in the MTPC region will also be briefly covered. Lastly, the performance of
the CW-IPM machine will be compared with two other similar-sized DW-IPM
machines.
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
171
7.2 CONSTRUCTION PROCESS
The construction process of the CW-IPM machine can be broken down into three steps:
the rotor and stator assembly; stator winding; and finally the total machine assembly.
The total manufacturing time of the final CW-IPM prototype was twelve months. This
section illustrates the abovementioned three steps of constructing the motor, as well as
the manufacturing timeline of the construction process.
7.2.1 Manufacturing Duration
The flow-chart in fig. 7.1 shows the entire construction process from the date when
designs were submitted to the manufacturers to the date the experimental setup was
completed. The reason for describing this process is to give future PhD
students/researchers a clearer picture of what has to be done if a prototype is to be
constructed. Key delays in the manufacturing process will be stated and suggestions to
speed up the process will be made.
Key delays for the manufacturing process and suggestions to avoid them include:
Delay: Redesigning of unmanufacturable portions of the initial model.
- Suggestion: Regular meetings with manufacturer during the design stage
Delay: Duration taken for the desired steel grade to be shipped.
- Suggestion: Choose and order desired core material before the design
Delay: Manufacturing errors in cutting of laminations.
- Suggestion: Not applicable.
Delay: Windings not being wound according to a specified layout.
- Suggestion: Commercial winders are not always adaptable to new winding
types. Coils can be fitted by the winders but phase connections can be done in
our labs.
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
172
Fig. 7.1 Duration breakdownand stepsof construction process
Meeting with CSIRO:
Discussion of design
Correction of
unmanufacturable
portions
Choice of alternate
steel grade
Ordering of magnets
from China (RareMag)
Correction of slot
opening size, magnet
slot size
Modelling performance
with available steel
grade
Odering of steel
(Sankey) and lazer
cutting (LazerXperts)
Rotor assembly (CSIRO):
Stacking laminations; insertion
of magnets; dynamic balancing
Stator assembly (CSIRO):
Stacking laminations; heat
shrinking into casing
Removal of original
stator and rotor
from ABB casing
Winding of stator
(Atom Electrical)
Assembly of machine
and experimental setup
(UNSW)
Start of
manufacturing
process
Month 1
to
Month 3
Month 4
to
Month 8
Month 9
to
Month 11
Month 12
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
173
7.2.2 Rotor Assembly
As mentioned in chapter 4, the material used for both the rotor and stator laminations
was 0.35mm thick (35RM300). A laser-cut rotor lamination of the prototype is shown in
fig. 7.2. The 79mm rotor stack was assembled by pressing one lamination at a time
down a shaft; tapered ends of the shaft simplify this process.
On completion of the stack, magnets are then inserted. Inserting the magnets is a fairly
easy task, as all the magnet pieces get drawn into the rotor stack. The completed rotor
stack with magnets inserted is shown in fig. 7.3.
Fig. 7.2 Laser-cut rotor lamination
It should also be noted that axially segmented magnets, (for the purpose of eddy current
loss reduction), can also be easily inserted piece by piece into the rotor. The segments
making up a magnet pole would not be repelled out of the slot as initially expected.
Magnet slots
4mm
d
holes
for bolts to
holding end-
plates and
laminations
together
6mm key hole
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
174
Fig. 7.3 Complete rotor stack with magnets inserted
Lastly the end plates are bolted on, bearings mounted and the rotor is dynamically
balanced. Dynamic balancing is done by drilling holes into the endplates to achieve
rotational weight balance.
Fig. 7.4 Completed dynamically-balanced rotor
Shaft
Magnets
Endplate
Bolts
Bearing
mounted
Holes
drilled in
endplate
to achieve
dynamic
balance
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
175
7.2.3 Stator Core Assembly and Stator Windings
The same rotor steel grade (35RM300) is used in the stator. The completed 80mm stator
stack is shown in fig. 7.5. Six equally spaced, w-shaped groves were cut into the outer
periphery of the stator for welding and alignment purposes. It should be noted that these
groves should be situated in the inter-slot section, so as not to hinder flux paths in the
stator yoke section.
Fig. 7.5 Completed stator stack
The stator was wound with the horizontal-fill method, as discussed previously in
chapter 4. The windings done by Atom Electrical are of acceptable quality, with 42%
slot-fill factor achieved. The initially achieved 14.5mm end winding length with the
plastic stator could not be achieved in the winding of the prototype stator; the actual end
winding length measured 18.5mm (shown in fig. 7.6).
Groves situated
in the inter-slot
sections
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
176
Fig. 7.6 Measurement of end winding length in UNSW CW-IPM stator
7.2.4 Overall Machine Assembly
Due to the high remanent flux density of the magnets, inserting the rotor into the stator
required the use of an aluminium sleeve. This method is tedious and causes damage to
both the stator and rotor laminations. An alternative method of assembling the machine
is to first remove the magnets from the rotor; insert the rotor; then re-insert the magnets.
For this machine, the time taken to assemble it using the latter method was slightly less
compared to the use of an aluminium sleeve. Fig. 7.7 shows the completed CW-IPM
machine assembly in comparison to the DW S-IPM machine.
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
177
Fig. 7.7 Comparison of UNSW CW-IPM machine assembly (right) and DW S-IPM machine
(left)
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
178
7.3 OPEN CIRCUIT PARAMETERS
In chapter 4, it was shown that with an 18-slot, 14-pole FE model, a high winding factor
and sinusoidal back EMF could be achieved. Due to the elimination of slot and pole
periodicity achieved by fractional-slot distribution, very low cogging torque can also be
achieved.
7.3.1 Back EMF
For back EMF measurements, the CW-IPM machine terminals were left open-circuited
and the machine was ran at various speeds by a prime mover (1kW Kollmorgen
machine). The measured line to line- and line to neutral- induced voltages were 154V
rms
and 89V
rms
respectively. These values were very similar to the modelled values, albeit
slightly lower (157V
rms
and 92V
rms
). The near-sinusoidal shape of the measured
waveforms also followed very closely to the modelled ones as shown in fig. 7.8.
Fig. 7.8 Measured line to line and line to neutral backEMF waveforms compared against
modelled values
-250
-200
-150
-100
-50
0
50
100
150
200
250
0.005 0.01 0.015 0.02 0.025
Modelled
Vab @50Hz
Measured
Vab @50Hz
Modelled
Van @50Hz
Measured
Van @50Hz
I
n
d
u
c
e
d

V
o
l
t
a
g
e

(
V
)
Time (s)
V
L-N(Measured)
=89V
rms
V
L-N(Modelled)
=92V
rms
V
L-L(Modelled)
=157V
rms
V
L-L(Measured)
=154V
rms
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
179
Fig. 7.9 shows the linear increase in back EMF and speed.
Fig. 7.9 Measured line to line back EMF versus speed compared against modelled values
7.3.2 Cogging Torque
Cogging torque in the CW-IPM machine is expected to have a high frequency carrier
signal with a low frequency modulating signal. The experimental setup shown in fig.
7.10 consists of a 0.6m beam, balanced on both sides, weights, and a position sensor. By
gradually adding weights at each position, the amount of torque required to rotate the
machine at various angles can be calculated. The cogging torque measurements made in
the clockwise direction make up the positive half of each fluctuation, and vice versa.
Fig. 7.10 Cogging torque measurement setup
0
100
200
300
400
500
600
700
800
0 500 1000 1500 2000
Modelled
Vab Vs.
Speed
Measured
Vab Vs.
Speed
Speed(rpm)
I
n
d
u
c
e
d

V
o
l
t
a
g
e

(
V
r
m
s
)
0.6m Beam
(Balanced on both sides)
Position Sensor
(Connected to DS1104 Board)
Weights (1g to 500g)
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
180
Fig 7.11a (measurements from 3-10) and 7.11b (measurements from 12-18.5)
compares the cogging torque points measured over 1 cogging torque period (20
mechanical degrees), with the results obtained from the FE model.
(a) Measurement range 3 to 10 degrees
(b) Measurement range 12 to 18.5 degrees
Fig. 7.11 Measured cogging torque points compared against results obtained from FE model
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
181
From fig. 7.11, it is shown that the measured points agrees with values from the FE
model, with slight deviations in the rotor angle. The peak magnitude of the measured
points were higher (approximately 2x higher) than the peaks of the waveform obtained
by the FE model. Considering that the cogging torque values are extremely small (<
0.018Nm
(p-p)
), effects on bearing striction and/or rotor eccentricity/balance (which
were not considered in the FE model), would contribute significantly to the magnitude
of measured torque points. Due to the high frequency fluctuations created by fractional
slot distribution, measured cogging torque points had to be taken in 0.2 degree intervals.
This contributed to the errors in the rotor positions where torque was measured.
From the measured cogging torque points, a curve achieved from FE analysis was fitted.
This curve termed as expected cogging torque waveform (shown in fig. 7.12), would
be used in place of the actual measured cogging torque waveform in the subsequent
sections as a comparison to the DW-IPM models.

Fig. 7.12 Curve fitted cogging torque waveform
It can be seen that the modulation of the cogging torque waveform is proportional to the
width of the slot (i.e. 2 cycles per slot), similar to that of regular integral slot machines.
Envelope for peak
cogging torque values
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
182
The only difference is the high frequency ripples, which are caused by the aperiodic
flux linkage in other slots and phases (a characteristic of fractional slot windings),
which has been reported in several recent papers as well.
7.3.3 Inductance and Saliency Ratio
The inductance waveform was measured using the AC standstill test method with
single-phase excitation, (stated in chapter 3). Static measurements were performed in
two degree increments throughout one electrical revolution, with the position being read
from a mounted encoder. The variation of voltage in self and mutual-phases was then
used to calculate the self- and mutual-inductance values at each position. The self and
mutual-inductance values (measured at 3A) from the finite element model, as shown in
chapter 4 are re-illustrated here as a comparison to the measured values. Fig. 7.13 shows
the inductance values from the FE model and fig. 7.14 shows the values measured from
the prototype.
Fig. 7.13 Modelled self- and mutual-inductance waveform from the FE model with 3A
rms
current excitation
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
183
Fig. 7.14 Measured self and mutual-inductance waveform from the prototype with 3A
rms
current
excitation
From fig. 7.13 and 7.14, it is seen that the average and peak to peak magnitude of the
self-inductance obtained in both the FE model and prototype machine are relatively
similar. While the mean value mutual-inductances are also relatively similar, the
variations in the prototype machine are much higher compared to the FE model. This
larger variation in mutualinductance leads to a higher saliency ratio of 1.12, compared
to an almost negligible saliency ratio of 1.04 in the FE model. L
d
and L
q
for the FE
model are 81.16mH and 84.45mH, whereas L
d
and L
q
for the actual machine are
82.52mH and 92.48mH respectively. The most probable reason for this difference is,
that 2D FE model was not able to account for some of the flux leakage occurring toward
the end of the windings. We can also conclude here that end-turn winding of the q-axis
inductance is more than the d-axis inductance. This is in fact advantageous for the
machine because it can be a tool to optimize the saliency ratio.
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
184
With different excitations, saturation of the core also causes the inductance to change
slightly. Fig. 7.15 compares the measured and modelled dq-axis inductance values
between 1A
rms
to 3A
rms
.
Fig. 7.15 Variation of dq-axis inductances with current
Fig. 7.15 shows that overall inductance values fall as the current increases. For the
measured inductance, the increase in current has a greater effect on the q-axis
inductance compared to the d-axis inductance, thus leading to a slight decrease in the
saliency ratio with higher currents.
0
20
40
60
80
100
1 1.5 2 2.5 3
I
n
d
u
c
t
a
n
c
e

(
m
H
)

Current (A)
d-axis inductance
(modelled)
q-axis inductance
(modelled)
d-axis inductance
(measured)
q-axis inductance
(measured)
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
185
7.4 STEADY STATE ANALYSIS
The steady state analysis of the machine includes the torque/power versus speed curves
as well as the i
d
and i
q
trajectory required to achieve the curves; the voltage trajectory
from zero to maximum speed; and efficiency over the CPSR. Steady state waveforms of
the input currents and voltages as well as i
d
and i
q
values will also be shown.
7.4.1 Torque and Power Characteristics
In Chapter 4, it is shown that a 10:1 CPSR can be achieved for the FE model with an
input line to line voltage of 240V
rms
and a base speed of 429rpm. Due to loading and
equipment constraints torque transducer, gearbox and loading generator are rated for a
shaft torque of approximately 10Nm as well as the desire to achieve a higher
efficiency, the voltage limit on the constructed CW-IPM prototype was increased:
Input line to line voltage: 340V
rms
Rated current: 2.2A
rms
Base speed with rated voltage: 573rpm
The main components of the experimental setup are shown in fig. 7.16. (the full setup is
shown in appendix F)
Fig. 7.16 Experimental setup to measure back EMF and torque/power versus speed performance
CW-IPM motor
(connected to inverter)
Torque
transducer
Loading generator
With external resistor
bank (not shown)
Gearbox
(4:1)
Encoder
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
186
The three-phase supply to the CW-IPM machine is from an inverter (diagram shown in
chapter 6, fig 6.7). The three-phase space vector modulation (SVM) generated supply
voltage and currents were derived based on controller outputs. The controller
implemented in C code, ran together with dSPACE control desk which provided the
graphical user interface on a Windows-based PC. Input and output signals of the
controller were handled by a DS1104 ADC/DAC control board.
From zero to base speed, the machine ran under MTPC conditions. For the CW-IPM
machine, it is assumed that under MTPC operation, i
d
is zero and i
q
is equal to the
supply current i
s
. This is due to the fact that the saliency ratio is close to unity. After
base speed, when the rated voltage is reached, i
d
is then made to oppose the flux
produced by the rotor magnets. This isdone to maintain constant voltage by suppressing
the induced back EMF as speed increases. The required i
d
and i
q
current trajectory to
maintain constant voltage with increasing speed is shown in fig. 7.17:
Fig. 7.17 Measured dq-axis current points under field weakening operation
Maximum FW point
MTPC
point
FW Region
Unstable region
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
187
The voltage measured throughout the speed range from operation in the MTPC region
to the maximum achievable speed in the field weakening region is shown in fig. 7.18:
Fig. 7.18 Measured line to line voltage versus speed
From the above figure, it can be seen that a constant voltage can be maintained after
base speed when the FW trajectory in fig. 7.17 is followed. Here the field-weakening
range/CPSR is taken from base speed (the point where the rated voltage is reached) to
the point where power drops below the value achieved at base speed. The torque and
power versus speed characteristics based on the current and voltage points from the
abovementioned figures are shown in fig. 7.19:
Fig. 7.19 Measured torque versus speedcharacteristics of the CW-IPM machine prototype
0
50
100
150
200
250
300
350
400
0 1000 2000 3000 4000 5000
Line to line voltage versus speed
I
n
p
u
t

v
o
l
t
a
g
e

(
V
L
-
L
)
2000 3000
Speed (rpm)
MTPC
Region
FW
Region
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
188
With a voltage limit of 340V
L-L
, greater than 7:1 CPSR (7.2:1 max) can be achieved.
Output power at base speed is 762W; over 800W of power is achieved just slightly
above base speed all the way to the 7:1 FWpoint. The maximum power of 905W was
achieved at 1290rpm. Over 80.8% efficiency was achieved from base speed till the 6.2:1
FW point. The efficiency of the machine versus speed is plotted in fig. 7.20:
Fig. 7.20 Measured efficiency versus speed
Within the 6.2:1 range (indicated in fig. 7.20) the efficiency varied from 80.8% to 83%.
Measured efficiency before and after this range dropped rapidly to 61.4% at the lowest
measured speed of 191rpm, and to 65.1% at the highest achieved speed of 4889rpm,
(which was the maximum speed where constant voltage can be maintained, 8.5:1 point).
Therefore, the optimal operating range in terms of efficiency would be between 573rpm
(base speed) and 3562rpm, resulting in a 6.2:1 CPSR. Comparison between the
experimental results and the predicted efficiency (by calculations and FE analysis)
showed result were very close (with 0.2 3% error). This justifies the loss predictions
shown in chapter 5.
0
10
20
30
40
50
60
70
80
90
100
0 500 1000 1500 2000 2500 3000 3500 4000
Measured Efficiency vs. Speed
Predicted Efficiency vs. Speed
>80%Efficiency over a 6.2:1 CPSR
E
f
f
i
c
i
e
n
c
y

(
%
)
Speed (rpm)
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
189
7.4.2 Steady State Voltage and Current Characteristics
This section illustrates the steady state speed, currents and voltages in the dq-axis
reference frame, as well as the three phase current and voltage input to the machine.
These readings were taken at base and maximum FW speed. The speed waveforms
under full load conditions are shown in fig. 7.21. The signals in red (#1:2) are the
desired controller references, and the signals in green (#1:1) are the following/actual
output signals.
(a) Base Speed (b) Maximum Speed
Fig. 7.21 Steady-state speed waveforms
The steady state speed output follows the reference with zero error at base and
maximum field-weakening speed. The following speed signal at maximum speed is
constant, but at base speed the signal has low magnitude perturbations. At these two
speeds current and voltage signals in their corresponding reference frames are shown in
fig 7.22.
Time(s)
(
R
P
M
)
Time(s)
(
R
P
M
)
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
190
(a) Base Speed (b) Maximum Speed
Fig. 7.22 Steady state i
d
current waveforms
Reference i
d
values are user inputs to the program and are therefore constant. The
following i
d
signals contain high frequency fluctuations of substantial magnitude. These
fluctuations, which are also present in the in the i
q
actual signal, lead to vibrations and
additional acoustic noise in the very low speed and near the maximum speed of the
machine.
(a) Base Speed (b) Maximum Speed
Fig. 7.23 Steady state i
q
current waveforms
The slight fluctuations in the reference i
q
current waveform were more pronounced at
higher speeds. The actual i
q
current waveforms follow the reference waveforms as long
as the required voltage does not exceed limited values. After the maximum speed of
Time(s)
(
A
)
Time(s)
(
A
)
Time(s)
(
A
)
Time(s)
(
A
)
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
191
4889rpm, i
q
becomes uncontrollable and develops following errors, thus the CW-IPM
drive is limited to this speed.
The corresponding current waveforms and voltage pulses in the abc reference frame for
steady state operation is shown in fig. 7.24 below: (Switching frequency used here is
10kHz).
(a) Current and voltage waveforms at base speed of 60rad/s
(b) Signals at maximum FW speed of 426rad/s
Fig. 7.24 Steady state current and voltage inputs to the machine in abc reference frame
Time (s)
C
u
r
r
e
n
t

(
A
)
V
o
l
t
a
g
e

(
V
)
V
o
l
t
a
g
e

(
V
)
Time (s)
C
u
r
r
e
n
t

(
A
)
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
192
7.5 TRANSIENT RESPONSE UNDER MTPCOPERATION
In this work, the steady state performances would be sufficient to verify the CPSR,
power density and efficiency of the prototype CW-IPM machine. Thus, only the basic
dynamic responsesin the MTPC region (standstill to base speed) will be shown. Control
strategies to improve the dynamic performance of the CW-IPM machine are currently
being studied.
7.5.1 Transient Voltage and Current Characteristics
Fig. 7.26 shows the speed step response of the CW-IPM drive:
(a) No load (b) Full load
Fig. 7.25 Speed step from standstill to base speed
(a) No load (b) Full load
Fig. 7.26 i
d
current waveforms with speed step from standstill to base speed
Time(s)
Time(s)
(
R
P
M
)
(
R
P
M
)
Time(s) Time(s)
(
R
P
M
)
(
R
P
M
)
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
193
The i
d
current waveforms show large negative spikes during the speed transition,
getting more pronounced with a higher load, despite the reference being zero.
(a) No load (b) Full load
Fig. 7.27 i
q
current waveforms with speed step from standstill to base speed
The actual i
q
current waveforms follow the reference well, however, high frequency
fluctuations are present in the reference signal.
The corresponding current waveform and voltage pulses for the speed step at no load
and full load in the abc reference frame is shown in fig. 7.28 below:
(a) No load
V
o
l
t
a
g
e

(
V
)
C
u
r
r
e
n
t

(
A
)
Time (s)
Time(s)
Time(s)
(
R
P
M
)
(
R
P
M
)
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
194
(b) Full load
Fig. 7.28 Current and voltage inputs to the machine in abc reference frame with step change in
speed
7.5.2 Torque Transients
The output torque of the machine is determined from a torque transducer mounted
between the drive and the loading machine. Fig. 7.29 shows the torque transient at full
load when the machine accelerates from standstill to base speed. Fig. 7.30 shows the
corresponding torque ripple when steady state is reached.
Fig. 7.29 Torque transient when CW-IPM machine accelerates from standstill to base speed at
full load
0
2
4
6
8
10
12
14
0 0.5 1 1.5 2
C
u
r
r
e
n
t

(
A
)
V
o
l
t
a
g
e

(
V
)
Time (s)
T
o
r
q
u
e

(
N
m
)
Time (s)
Section of torque
ripple shown in
fig. 7.31
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
195
Fig. 7.30 Measured torque ripple at steady state
The output torque ripple was 2.1Nm
(peak to peak)
, which amounts to 16.5% of the total
torque produced at base speed. The magnitude of torque ripple is higher than predicted.
This could be due three main reasons:
i) The noisy gearbox (in terms of vibrations) at the load resulting in relative
fluctuations at the torque transducer.
ii) The second reason being the controller design. More detailed parameter
identification and control strategies which are currently being implemented
fall beyond the scope of this thesis. This strategies, implemented on the
CW-IPM machine will be shown in future publications.
iii) Lastly, increased torque ripple could be caused by the load machine and
misalignment in coupling.
0
2
4
6
8
10
12
14
1.8 1.85 1.9 1.95 2
Torque ripple magnitude =2.1Nm
(peak to peak)

T
o
r
q
u
e

(
N
m
)
Time (s)
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
196
7.6 PERFORMANCE COMPARED TO DISTRIBUTED WINDING IPM
MACHINES
The main aim of this thesis was to design and build a CW-IPM machine to achieve
higher torque/power density, wider CPSR, lower cogging torque, and of equivalent
efficiency as two other available DW-IPM machines of equal size [13]. As a
comparison, two other previously constructed, equally sized DW-IPM machines will be
used. All three machines were designed to fit in the same 550W ABB casing. The two
DW machines have 4-poles, while the CW machine has 14-poles. This is due to the
required 14-pole, 18-slot layout to achieve low cogging torque and an appropriate back
EMF waveform, as mentioned in earlier chapters. Since pole numbers are not equal, a
comparison of output torque would not be appropriate. Most other quantities such as
CPSR, output power, cogging torque and efficiency can still be compared.
The first DW IPM machine (shown in fig. 7.31a) has regular single-pieced, flat shaped
poles. It uses sintered magnets with B
r
=1.05T. This machine will be named IPM-I.
The second DW-IPM machine, (shown in fig. 7.31b), has segmented magnets. It uses
bonded magnets with B
r
=0.78T. This machine will be named S-IPM.
It should be pointed out that the IPM-I was designed based on prior knowledge and
experience. No optimisation strategy was implemented. The S-IPM machine was
designed and optimised in a similar fashion to the CW-IPM machine- with FE analysis
as shown in this work. Additional details on its optimisation can be seen in [174].
The final CW-IPM machine model is shown in fig. 7.31c. It uses a similar magnet grade
as IPM-I with B
r
=1.04T.
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
197
(a) Single piece per pole IPM with DW (DW IPM-I)
(b) Segmented IPM machine with DW (DW S-IPM)
(c) Constructed CW-IPM prototype(CW-IPM)
Fig. 7.31 Comparison of three UNSW IPM machines
DW S-IPM
DW IPM-I
CW-IPM
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
198
7.6.1 Power and Torque versus Frequency Comparison
As the CW-IPM machine has 14 poles and the two other DW-IPM machines have only
4 poles each, frequency is used as the basis of comparison instead of mechanical speed.
Fig. 7.32 shows the power versus frequency comparison between the three IPM
machines. It should be noted that whilethe comparison against the DW IPM-I would
not be fair due to the fact that the machine was not optimised for field weakening
performance nor for high power density. It is included to show that despite a much
lower saliency ratio, magnet volume and magnet energy density, the design is still the
most crucial inachievinggood field weakening performance.
Fig. 7.32 CPSR comparison between the three UNSW IPM machines
From the results obtained, it is shown that the CW-IPM machine not only achieved a
much wider CPSR >6.2:1, compared to 4:1 in the S-IPM machine and almost no field
weakening capability in IPM-I it also achieved a 56% power increase over the
constant power region as compared to the two other DW-IPM machines.
In terms of peak torque under maximum torque per unit current (MTPC) operation, the
CW-IPM machine achieved a peak shaft torque of 12.7Nm, compared to 2.25 Nm and
0
100
200
300
400
500
600
700
800
900
1000
0 50 100 150 200 250 300 350 400 450
P
o
w
e
r

(
W
)

Frequency (Hz)
>6.2:1 CPSR
4:1 CPSR
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
199
2.3Nm in the S-IPM and IPM-I respectively. For a fair comparison of torque between
the 14-pole CW-IPM machine and the 4-pole DW-IPM machines, torque is normalised.
The torque magnitude of 1p.u. is assumed in the MTPC region.
Fig. 7.33 Normalised output torque comparison between the three UNSW IPM machines
The torque versus frequency waveforms clearly show that IPM-I has almost no field
weakening capability, with torque rapidly falling at the beginning of the field
weakening region. On the other hand, it is shown that both the S-IPM and CW-IPM
have excellent field weakening capability.
7.6.2 Cogging Torque Comparison
Comparing the two DW machines the S-IPM machine uses magnets with a lower
remanent flux density of 0.78T as compared to IPM-I, which uses magnets with remanet
flux density of 1.05T. Thus the cogging torque is naturally higher in the latter. However,
despite the useof high remanent flux density magnets in the CW-IPM machine (1.04T),
the cogging torque is substantially lower compared to the DW machines due to the
elimination of periodicity of slots and poles by using fraction-slot distribution.
0
0.2
0.4
0.6
0.8
1
0 50 100 150 200 250 300 350 400 450
T
o
r
q
u
e

(
P
.
U
)

Frequency (Hz)
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
200
Cogging torque comparisons between the three IPM machines is shown in fig. 7.34:
Fig. 7.34 Cogging torque comparison between the three UNSW IPM machines
From the comparison of the cogging torque values, it is shown that the CW-IPM
machine produced lower cogging torque magnitude, compared to the two other DW-
IPM machines. As expected IPM-I produced the largest peak cogging torque magnitude.
As higher generated torque will make the effects of cogging torque less significant,
another important comparison is, the amount cogging torque produced as a percentage
of the total torque generated in the MTPA region. This comparison is shown in fig.7.35
as follows:
Fig. 7.35 Cogging torque as a percentage of output torque at base speed comparison between
the three UNSW IPM machines
0.038Nm(p-p)
0.126Nm(p-p)
0.502Nm(p-p)
0.3%(p-p)
8.1%(p-p)
24.2%(p-p)
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
201
7.6.3 Efficiency Comparison
Data on efficiency versus speed for IPM-I is not readily available, although it is known
to be somewhere in the 60-65% range at rated speed, so efficiency was compared only
between the CW-IPM and SEG-IPM as shown in fig. 7.36:
Fig. 7.36 Efficiency comparison between the CW-IPM and S-IPM machine up to 200Hz
The SEG-IPM machine produced 84 to 85% efficiency from base speed up to the
measured 200Hz. In comparison the CW-IPM machine produced 80.8 to 83% efficiency
from base speed up to 420Hz. Copper loss, which is 2.4 times lower in the S-IPM
machine as compared to the CW-IPM machine, was the main reason for the lower
efficiency of the CW-IPM machine. Besides copper loss, other losses were lower in the
CW-IPM machine, despite the increase in MMF harmonics created by CW.
0
10
20
30
40
50
60
70
80
90
100
0 50 100 150 200
CW-IPM SEG-IPM Poly. (CW-IPM)
E
f
f
i
c
i
e
n
c
y

(
%
)

Frequency (Hz)
Base speed
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
202
7.6.4 Magnet Volume Comparison
Due to increasing magnet prices, the amount of magnet used for present-day machine
designs is subjected to constraints. Thus it is of interest to briefly compare the amount
of magnet used in the three machines. Fig. 7.37 compares the volume of magnet per kW
in each machine.
Fig. 7.37 Comparison of Magnet volume per kW the three UNSW IPM machines
This factor can be further improved if the efficiency of the machine is optimised. It
should be noted that, despite similar amount of magnet material used in the CW-IPM
and SEG_IPM, a magnet grade with higher energy density was chosen to compensate
for the loss in reluctance torque. This was done at the expense of higher rotor core
saturation and higher magnet losses.
Chapter 7: Construction and Performance Analysis of the Concentrated Winding IPM Machine Prototype
203
7.7 CONCLUSION
This chapter firstly illustrated the manufacturing process of the prototype CW-IPM
machine. Problems were identified and manufacturing delays were stated as a guide to
improve the manufacturing process of future machines.
The measured performance characteristics of the constructed CW-IPM prototype were
shown and used to verify the performance of the FE model. It was shown that the
measured results from the prototype agreed with the FE results with a high degree of
accuracy. The machine parameters and performance characteristics from the FE model
were compared to two other equally sized DW-IPM machines. These results showed
that the CW-IPM outperformed the two other DW-IPM models in terms of power
density (up to 56% higher), CPSR (greater than 55%) and cogging torque performance
(almost negligible compared to the other two DW machines when compared as a
percentage of total torque). Despite the high copper loss, an 80.8% efficiency was
achieved throughout a 6.2:1 CPSR.
This chapter has verified studies and the design of the CW-IPM machine in this work.
The successful design of the first CW-IPM prototype in our labs has provided a strong
basis and confidence in the FE models.
204
CHAPTER 8
EFFICIENCY OPTIMISATION AND SCALABILITY OF THE
CONCENTRATED WINDING IPMMACHINE
8.1 INTRODUCTION
In previous chapters, the design and experimental verification of the CW-IPM machine
have been shown. The close correlation of the experimental and FE results gave
confidence for further designs and optimisation. On the basis of the original 800W FE
model, the scalability and efficiency optimization of the CW-IPM machine will be
studied in this chapter.
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
205
8.2 EFFICIENCY OPTIMISATION
In chapter 7, it was shown that the CW-IPM machine operated at over 80% efficiency
throughout the CPSR. It was also shown, in chapter 5, that the majority of losses (over
90%of losses) in the CW-IPM machine weredue to copper loss. With the main focus
on minimising copper loss, winding modifications and two geometric changes to the
original design are proposed to increase the efficiency of the machine.
8.2.1 Winding Modification
Two hand-winding methods were introduced in chapter 4 the vertical-fill method and
the horizontal-fill method. The latter method was used in the prototype to reduce the
time and cost required to wind the machine. Here, for the purpose of optimising
efficiency, the vertical slot-fill method is used despite the increase in time and cost of
winding the machine. A 45% slot-fill factor was personally achieved in the prototype
stator. Thus a 45% slot-fill factor will be assumed in this optimisation study. The
difference in winding span and end winding length between these two methods is shown
in fig. 8.1 below:
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
206
(a) Horizontal slot-fill factor
(b) Vertical slot-fill factor
Fig. 8.1 Axial length comparison between two different slot-fill methods
With the vertical winding method, the overall length per turn is reduced significantly
(from 276mm per turn to 219mm per turn).
8.2.2 Proposed Designs for Efficiency Optimisation
The basic structure of the original design and the volume of the machine are preserved
in this modifications. For convenience, the first modified design will be called CW-IPM
R. For the CW IPM-R, the outer diameter and stack length of the machine ispreserved
(fig 8.2a). The rotor diameter of the machine is made smaller to increase the space in the
stator for larger slots, permitting the use of larger conductors, thus lowering copper loss.
The second design will be called CW-IPM S. For the CW-IPM S, the outer diameter is
increased but the stack length is reduced by half in order to keep the volume constant
117mm axial winding length
(measured)
21mm average winding span
95mm axial winding length (expected)
14.5mm average winding span
Horizontal slot-fill
Vertical slot-fill
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
207
(fig. 8.2b). The performance constraints are to preserve the same 6:1 CPSR and power
must be _ 800W over the CPSR.
(a) (b)
Fig. 8.2 Designs used for efficiency optimisation indicating outer dimensions
Table 8.1 below, gives key parameters of the two optimised designs. Fig. 8.3a and 8.3b
give performance characteristics for CW-IPM R and CW-IPM S respectively.
(a)
80mm
1
8
4
m
m
1
3
0
m
m
Volume =
1063mm
2
(b)
40mm
Volume =
1063mm
2
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
208
Table 8.1
Key Specifications of the Two Optimised CW-IPM Machine Designs
CW-IPM R CW-IPM S
Stator outer diameter 130mm 183.85mm
Rotor outer diameter 36mm 45.7
Airgap length 1mm 1.41mm
Slot opening width 1.2mm 1.2mm
Stack length 80mm 40mm
Rated current 2.2A
rms
/ph 2.2A
rms
/ph
Current density 4.25x10
6
A/m
2
5.37x10
6
A/m
2
Conductor size AWG 20 AWG 21
No. of turns per coil 163turns 127turns
Stator resistance 4.5O/ph 5.5O/ph
Mag. remanent flux 1.13T 1.13T
Slot-fill factor 45% 45%
(a) CW-IPM R
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
209
(b) CW-IPM S
Fig. 8.3 Efficiency and power versus speed performance of the two efficiency optimised model
The results firstly show a significant increase in efficiency, thus increasing in output
power for both optimised designs compared to the original design (where 80%
efficiency was achieved). Comparing the two optimised designs, the CW-IPM S
achieves a higher efficiency of 93%, while the CW-IPM R achieves an efficiency of
91% throughout the CPSR. Despite a higher stator resistance, the CW-IPM S achieved a
higher efficiency, due to its capability of producing higher maximum output power. The
maximum output power in the CW-IPM R was limited by the stator and rotor outer
radius.
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
210
8.3 SCALABILITY OF THE CONCENTRATED WOUND IPMMACHINE
This scalability study will be divided into two steps: Firstly, the machine parameters of
the original 800W prototype will be varied to study their effects on performance. From
this, a set of general design rules will be derived to ensure that the material properties
are fully utilised and that the machine is operating with the desired field-weakening
capability. Secondly, two up-scaled versions of the 800W prototype will becreated. By
implementing the design rules, it will be shown that the predicted power density and
efficiency of the scaled version can be increased significantly. FE analysis will be used
as the basis of this optimisation.
8.3.1 Airgap Length Variation
CW results in increased harmonic and sub-harmonic content in the MMF waveform,
causing localised saturation, thus, affecting the field-weakening performance of the
machine, (as explained in chapter 4). The application of double-layer stator windings as
well as a larger airgap helps lower the effects of these harmonics. From the original
800W, CW-IPM design, the effects of varying airgap length on machine performance
are investigated.
Fig. 8.4 shows the effects on the CPSR and input power, (at base speed), when airgap
length is varied from 0.6 to 1.6mm.
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
211
Fig. 8.4. Airgap length versus CPSR and input power
Fig. 8.5 shows the effects on total core loss, (measured at 500Hz), and overall efficiency
with the same variations in airgap length.
Fig. 8.5 Airgap length versus efficiency and core loss
From fig. 8.4, it can be seen that there is an obvious trade-off between input power and
the maximum achievable CPSR of the machine. While power falls almost linearly as
airgap length is increased, the CPSR increases exponentially with airgap length.
Fig. 8.5 shows that having a wider airgap length is beneficial in decreasing overall core
losses. However as airgap length increases, input power naturally decreases as well.
This makes the copper loss more prominent, thus reducing efficiency.
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
212
8.3.2 Magnet Strength and Armature Current Variation
In order to achieve optimal field weakening range, both the saliency ratio and
characteristic current values need to be considered. It was explained in chapter 3 that
saliency ratio optimisation was not effective in the CW-IPM machine. Thus,
optimisation of CPSR would be based solely on optimising the characteristic current
condition.
An increase in magnet strength results in an increase in characteristic current, (reiterated
in (8.1)). Therefore when magnets of higher remanent flux density are chosen, the
machine has to be designed with a higher rated current.
I
u
= I
ch
=

pm
I
d

(8.1)
The geometry of the machine was not altered; hence L
d
is kept the same (81.16mH) as
before. Fig. 8.6 shows the variation of magnet remanent flux density and the current
required to satisfy the condition (8.1) and achieve a >6:1 CPSR. Fig. 8.6 also shows the
lossless power produced at each point. The same core material specifications used for
the prototype CW-IPM machine 35RM300 (with a saturation magnetization of 1.68T at
5000A/m) is used here.
Fig. 8.6 Magnet remanent flux density versus input current and input power
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
213
It can be seen from fig. 8.6 that the current required to achieve equilibrium conditions
and the input power of the machine increase linearly with magnet remanent flux density,
until the core begins to saturate, (after I
s
=3.3A/B
r
=1.32T). Fig. 8.7 shows the flux
density plot for the machine under unsaturated, (with 2.3A
rms
of excitation current), and
saturated conditions, (with 4.1A
rms
of excitation current).In order to fully exploit the
material properties of the core, the machine should be made to operate at this
equilibrium point (I
s
=3.3A/B
r
=1.32T equilibrium is different or eachdesign).
(a) Unsaturated conditions (b) Saturated conditions
Fig. 8.7. Flux density plot of CW-IPM machine under saturated and unsaturated conditions
As magnet remanent flux and stator current are increased, the core loss increases almost
proportionately, while the copper loss increases with the current squared. It is of interest
to know the extent to which losses and efficiency are affected as the power density of
the machine increases with magnet remanent flux density. Fig. 8.8 shows the total
losses and overall efficiency of the machine as magnet remanet flux and current are
increased.
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
214
Fig. 8.8 Magnet remanent flux density versus total machine losses and efficiency
This section shows the effects of varying parameters in particular, airgap length,
magnet remnant flux density and rated current of the CW-IPM prototype machine
without significant alterations to the machine geometry. The study done here provides
designers with a set of rules to design the CW-IPM for field weakening operation.
These steps are listed as follows:
i. Decide on the desired split-ratio (the ratio of stator inner diameter to outer diameter).
A larger split-ratio would result in higher torque densities but decreased space for
stator slots, hence higher copper loss.
ii. Decide on the airgap length of the machine. A smaller airgap would result in higher
torque density and higher efficiency but a narrower CPSR.
iii. Decide on the required slot area, keeping the core saturation limits in mind. From the
FE model, ensure that the maximum flux density in the tooth and yoke are similar.
iv. Vary slot-opening width to fit desired conductor sizes. Essentially, slot-opening
widths should be kept as small as possible to achieve maximumflux linkage across
the airgap.
v. Determine the optimal equilibrium point of the machine (the point after which current
increases non-linearly with magnet remanent flux density to maintain desired CPSR).
Operating after this point yields a non-linear decrease in efficiency.
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
215
8.3.3 Effects of Scaling the Machine Size
The purpose of this section is to study the effects of scaling up the CW-IPM machine
size, as well as to observe the effects of applying the abovementioned design rules to
larger machines. The aim is to achieve over 6:1 CPSR, over 80% efficiency throughout
the CPSR, as well as to fully utilise the steel properties, while operating at the optimal
equilibrium point.
Two designs are created based on the proven 800W model. In the first model, only the
machine outer diameter is increased by a factor of two; the stack length remains the
same as the 800W machine (80mm). In the second model, the outer diameter is three
times that of the 800W machine and the stack length is also increased by two (160mm)
as shown in fig. 8.9.
Fig. 8.9. Comparison between the three machine sizes
The output torque (I
out
) on machine sizing can be determined by (8.2), where the
output torque/power of the machine increases linearly with the effective stack length
(l
c]]
) of the machine, and with the square of the machine outer diameter (
so
).
Size of the
800W
prototype
1
st
model
2
nd
model
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
216
I
out
= K
t

so
2
l
c]]

(8.2)
Assuming the torque constant K
t
is unaltered and the entire machine is scaled up
proportionately, then the output power for the first and second scaled models should be:
1st model: P
out
= I
(800w)


so(1st)
2

so(800w)
2

m

P
out
= 13.33
0.26
2
0.13
2
60 = 3. 2kW
2nd model: P
out
= I
(800w)


so(2nd)
2

so(800w)
2

l
c]](2nd)
2
l
c]](800w)
2

m

P
out
= 13.33
0.39
2
0.13
2

0.16
0.08
60 = 14. 4kW
where,
I
(800w)
=Torque of the 800W prototype machine at base speed

m
=Base speed of the machine

so(800w)
=Diameter of the 800W prototype machine

so(1st)
=Diameter of the 1
st
scaled model (refer to fig. 8.9)

so(2nd)
=Diameter of the 2
st
scaled model (refer to fig. 8.9)
l
c]](800w)
=Stack length of the 800W prototype
l
c]](2nd)
=Stack length of the 2
st
scaled model (refer to fig. 8.9)
It is obvious that proportional scaling of the machine will not necessarily yield the same
CPSR. Also, the maximum flux density in the stator core will decrease as the machine
outer diameter is increased, leading to under-utilisation of the core material. Thus, the
design rules mentioned in the previous section are applied to the scaled models with the
aim of achieving at least a 6.2:1 CPSR, greater than 80% efficiency and increased
torque/power density by operating at the optimal equilibrium point of the core material.
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
217
By following the design steps, 5kW and 30kW can be obtained from the two models,
(3.2kW and 14.4kW), respectively. Table 8.2 shows the design specifications of the
original 800W CW-IPM prototype, in comparison to the optimised 5kW and 30kW
machines.
Table 8.2
Key Specifications of the Three CW-IPM Machine Designs
0.8kW
CW-IPM
5kW
CW-IPM
30kW
CW-IPM
Stator outer diameter 0.13m 0.26m 0.39m
Rotor outer diameter 0.080m 0.160m 0.240m
Airgap length 1.2mm 3.0mm 3.6mm
Slot opening width 1.2mm 3.0mm 4.8mm
Stack length 0.080m 0.080m 0.160m
Rated voltage 320V
rms(l-l)
320V
rms(l-l)
320V
rms(l-l)
Rated current 2.2A
rms
/ph 13.9A
rms
/ph 75.2A
rms
/ph
Conductor size AWG 22 AWG 13 AWG 5
No. of turns per coil 115turns 53turns 16turns
Mag. remanent flux 1.04T 1.34T 1.36T
Core saturation mag. 1.67T
@5000A/m
1.67T
@5000A/m
1.67T
@5000A/m
Slot fill factor 41% 41% 41%
The input and output power versus speed characteristics of the 5kW and 30kW
machines are compared in fig. 8.10 and 8.11 respectively:
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
218
Fig. 8.10 Input power, output power and efficiency versus speed characteristics of the 5kW
design
Fig. 8.11 Input power, output power and efficiency versus speed characteristics of the 30kW
design
The results show that the scaled designs were both able to achieve over 8:1 CPSR. The
5kW model achieved over 92.2% efficiency and 30kW model achieved over 95.8%
efficiency throughout the CPSR.
In the optimised designs, the core material operates with higher flux densities, thus core
loss would naturally increase. Despite the increase in core loss, efficiency is increased
due to the significant reduction of copper loss (which contributes to the largest portion
of losses in the machine) with a larger machine outer diameter. Fig. 8.12 shows the loss
breakdown of the three machine models:
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
219
Fig. 8.12. Losses in the three machine sizes as a percentage of total loss
From scaling the size of the machine, the following key points can be noted:
While increasing the machine outer diameter results in an exponential increase
in power, increasing the stack length results only in a linear increase in power, in
compliance to (8.2). Additionally, increasing the machine outer diameter
increases the efficiency of the machine.
Changing the conductor size to achieve the desired rated voltage has no effect on
copper loss and output power, as long as ampere-turns and the slot-fill factor are
unchanged.
Chapter 8: Efficiency Optimisation and Scalability of the Concentrated Winding IPM Machine
220
8.4 CONCLUSION
This chapter has shown that the original 800W FE model can be effectively optimised
to provide high efficiency (over 93% throughout the CPSR), by the reduction of copper
loss. The comparison of two optimized models with the same volume, shows that
having a larger outer diameter and shorter stack length was effective in increasing both
efficiency and torque density of the machine.
The scalability of the machine was also studied in this chapter. Through variation of
machine parameters, a general set of design rules was illustrated to fully utilise material
properties as well as to achieve desired CPSR and power density. By applying these
design rules on two scaled models, power density in the 3.2kW and 14.4kW model can
be optimised to produce 5kW and 30kW respectively. Efficiency of over 92.2% and
95.8% were achieved by each model respectively.
221
CHAPTER 9
CONCLUSION AND SUGGESTIONS FOR FUTURE WORK
9.1 CONCLUSION
The work done in this thesis has proven that the fractional-slot CW-IPM machine is a
suitable candidate for field-weakening applications. The constructed double-layer CW-
IPM machine with 14 v-shaped poles and 18-slots showed that a very wide 7.2:1 CPSR
could be achieved. Efficiency of over 80% was achieved over a 6.2:1 CPSR. The CW-
IPM also achieved a higher torque density, and much lower cogging torque compared to
two other equally sized DW-IPM machines.
In chapter 3, open-circuit characteristics of the CW-IPM machine were studied. Several
slot and pole combinations were investigated, and it was shown that the 18-slot, 14-pole
model achieved the highest induced back EMF magnitude, as compared to other 14-pole,
fractional-slot combinations. The calculated winding factor of 0.902 was confirmed by
comparison with an equivalent integral-slot DW machine model. This combination of
slots and poles also resulted in a near-perfectly sinusoidal EMF waveform, and very low
cogging torque magnitude. The comparison of two CW-IPM machines with different
magnet geometries, (rectangular and v-shaped), with the CW-SPM machine, showed
that the machine with rectangular, single-piece/pole magnets was inferior to the CW-
SPM machine interms of torque density, and field-weakening capability. The CW-IPM
machine with v-shaped magnets, on the other hand, had improved field weakening
capability, but with slightly lower peak torque density as compared to the CW-SPM
machine, due to saturationeffects.
Chapter 9: Conclusion and Suggestions for Future Work
222
In chapter 4, the CPSR of the 14-pole, 18-slot, v-shaped IPM design was optimised by
v-angle and airgap variations. It was shown that by satisfying the characteristic current
equilibrium conditions by variation of specific machine, the optimal CPSR could be
achieved. From optimisation done in this chapter, the final design was specified.
Parameters and performance characteristics of the final model were determined by FE
analysis.
In chapter 5, a detailed study of losses in the CW-IPM machine was done. It showed
how the increase in harmonics resulting from CW affected frequency related losses.
Electromagnetic losses, (consisting of core, magnet and I
2
R losses), and mechanical
losses, (consisting of bearing and windage losses), were studied and quantified. This
study highlighted the importance of choosing thin silicon steel laminations for the rotor
and stator core material. It also indicated that magnet losses in the CW-IPM were very
low, and that loss reduction by magnet segmentation was not as effective as magnet
segmentation in the equivalent CW-SPM machine. It was shown that the calculated
mechanical losses were low even at maximum operating speed of the machine. Chapter
5 also showed that copper loss made up for the majority of loss (>90%) inthe machine;
if the efficiency of the machine were to be improved, the main focus should be on
minimising copper loss.
Chapter 6 presented the control methodology of the CW-IPM machine. It showed the
voltage and current limits that the drive is subjectedto by use of the circle diagrams. It
was shown that the widely-used trajectories calculated by Morimotos equations led to
the over-weakening of the magnet fields. Thus a manually obtained i
d
current
trajectory was used in the control of the prototype CW-IPM machine.
Chapter 9: Conclusion and Suggestions for Future Work
223
Chapter 7 firstly illustrated the manufacturing process of the prototype CW-IPM
machine. Problems identified and manufacturing delays were stated as a guide to
improve the manufacturing process of future machines. The measured performance
characteristics of the constructed CW-IPM prototype was shown and used to verify the
performance of the FE model. It was shown that the measured results from the
prototype agreed with the FE results with a high degree of accuracy. The machine
parameters and performance characteristics from the FE model were compared to two
other equally sized DW-IPM machines. This comparison showed that the CW-IPM
outperformed the two other DW-IPM models in terms of power density, (up to 56%
higher), CPSR, (7.2:1 as opposed to a maximum of 4:1 achieved by the DW S-IPM), as
well as cogging torque performance, (0.3%
(p-p)
as compared 8.1%
(p-p)
and 24.2%
(p-p)
achieved by the other two machines as a percentage of total torque). Efficiency was
slightly lower in the CW-IPM machine, (80.8 to 83% over a 6.2: CPSR), as compared
with the DW S-IPM machine, (84 to 85% over a 4:1 CPSR).
With the successful design of the CW-IPM, achieving desired performance
characteristics, as well as the confidence gained from the close correlation of FE and
measured results, an efficiency optimization and scalability study was performed.
Chapter 8 showed that the original 800W FE model could be effectively optimised to
provide high efficiency, (up to 93%). Two optimised models with the same volume
were compared. It was shown that the model with larger outer diameter and shorter
stack length achieved a higher efficiency, as compared to one with the same dimensions
as the prototype machine, but with a smaller rotor. Chapter 8 also studied the scalability
of the machine. Through variation of machine parameters, a general set of design rules
were illustrated. This rules aid in designing the CW-IPM machine to achieve the desired
Chapter 9: Conclusion and Suggestions for Future Work
224
CPSR, and power density, as well as to fully utilise material properties. Two scaled up
models of the CW-IPM machine were created. Based on theoretical calculations, the
presumed output power of the two scaled models were 3.2kW and 14.4kW. However,
by application of the abovementioned design rules, the optimised models each produced
5kW and 30kW respectively.
The successful design, construction of the first CW prototype in our labs has provided a
strong basis, and confidence, in the field-weakening capability of the CW-IPM machine.
It has also created opportunities for future work to be done in this area, such as
performance optimisation, control implementation, as well as to determine the
suitability of the CW-IPM for various industrial applications.
Chapter 9: Conclusion and Suggestions for Future Work
225
9.2 SUGGESTION FOR FUTURE WORK
Due to the lack of a suitable loading generator, (capable of producing required torque
over the entire speed range), sensing equipment (torque transducer), as well as a time
constrain, the torque and field-weakening capability of the machine could not tested to
its maximum limits. With further testing using appropriate equipment, increased power
and a wider CPSR can be achieved. This work will soon be carried by a future student,
when the equipment becomes available. From simulations, it is noted that the saturation
in the steel is low (typically lower than 1.6T). This indicates that the tooth width and
yoke length can be made smaller to provide space for increased winding size- hence
lower copper loss.
In this thesis, it was shown that commonly used vector control techniques could not be
applied to the CW-IPM machine, thus a manually obtained i
d
and i
q
trajectory was used
to operate the machine at optimal power throughout the speed range. With this method,
dynamic performance of this machine could not be fully tested. Therefore, a proper
control technique has to be implemented to achieve desired dynamic performance.
Subsequently, sensorless control can also be implemented. Above base speed (60rads),
common DTC control methods should work. However, due to the low saliency ratio of
1.12, achieved by the CW-IPM machine, the closed-loop observer and high-frequency
signal injection methods might not work. If a higher saliency ratio is required, the
machine can be redesigned based on rotor magnet geometry variation techniques
illustrated in appendix B, as well as the investigation of end-winding effects,
(contributing the q-axis inductance), can be done to further increase the saliency ratio.
Chapter 9: Conclusion and Suggestions for Future Work
226
Additional focus can be placed upon trying to improve the torque/power density of the
CW-IPM machine. Also, as the cost of rare earth magnets escalating, it would be
beneficial to design machines with less magnet material (NdFeB in particular).
Lastly, with a high number of poles, as well as the capability of achieving high
efficiencies, CW-IPM may by suitable for low speed applications, such as in a wind
generator.
227
REFERENCES
[1] H. W. Beaty and J . L. Kirtley, Electric motor handbook, 1998.
[2] J . M. Miller, Propulsion Systems for Hybrid Vehicles. United Kingdom: The
Institution of Engineering and Technology, 2004.
[3] D. Hans[man, Brushless Permanent Magnet Motor Design, 2nd ed.: The
Writers' Collective, 2003.
[4] Z. Q. Zhu and D. Howe, "Halbach permanent magnet machines and applications:
a review," Electric Power Applications, IEE Proceedings -, vol. 148, pp. 299-
308, 2001.
[5] A. M. El-Refaie, "High Speed Operation of Permanent Magnet Machines," PhD
Thesis PhD, University of Wisconsin-Madsion, 2005.
[6] T. Yamakawa, S. Wakao, K. Kondo, and T. Yoneyama, "A new flux weakening
operation of interior permanent magnet synchronous motors for railway vehicle
traction," in European Conference on Power Electronics and Applications 2005,
pp. 1-6.
[7] N. Bianchi, S. Bolognani, and B. Ruzojcic, "Design of a 1000 HP Permanent
Magnet synchronous motor for ship propulsion," in 13th European Conference
on Power Electronics and Applications, EPE, 2009, pp. 1-8.
[8] T. M. J ahns and R. C. Van Nocker, "High-performance EHA controls using an
interior permanent magnet motor," Aerospace and Electronic Systems, IEEE
Transactions on, vol. 26, pp. 534-542, 1990.
[9] S. Suzuki, T. Katane, H. Saotome, and O. Saito, "Electric power-generating
system using magnetic coupling for deeply implanted medical electronic
devices," Magnetics, IEEE Transactions on, vol. 38, pp. 3006-3008, 2002.
[10] J . Zhang, Z. Chen, and M. Cheng, "Design and comparison of a novel stator
interior permanent magnet generator for direct-drive wind turbines," Renewable
Power Generation, IET, vol. 1, pp. 203-210, 2007.
[11] P. Kyoung-J in, J . Dae-Sung, C. Hong-Soon, B. J ae-Nam, and L. J u, "A study on
the reducing of cogging torque of IPMSM," in International Conference on
Electrical Machines and Systems, ICEMS, 2008, pp. 3137-3141.
[12] W. L. Soong and N. Ertugrul, "Field-weakening performance of interior
permanent-magnet motors," Industry Applications, IEEE Transactions on, vol.
38, pp. 1251-1258, 2002.
[13] R. Dutta and M. F. Rahman, "Design and Analysis of an Interior Permanent
Magnet (IPM) Machine With Very Wide Constant Power Operation Range,"
Energy Conversion, IEEE Transaction on, vol. 23, pp. 25-33, 2008.
[14] Toyota, "Toyota hybrid system - THS II," Toyota Motor Corporation, Public
Affairs Division, Tokyo, J apan2003.
[15] L. Chong, R. Dutta, and M. F. Rahman, "Electromagnetic Losses in a 1kW
Concentric Wound IPM Machine for Field Weakening Applications," Journal of
Applied Superconductivity and Electromagnetism, vol. 1, pp. 95-100, 2010.
[16] J . Cros and P. Viarouge, "Synthesis of high performance PM motors with
concentrated windings," Energy Conversion, IEEE Transaction on, vol. 17, pp.
248-253, 2002.
References
228
[17] F. Magnussen and C. Sadarangani, "Winding factors and J oule losses of
permanent magnet machines with concentrated windings," in Electric Machines
and Drives Conference, IEMDC'03. IEEE International, 2003, pp. 333-339
vol.1.
[18] A. M. El-Refaie and T. M. J ahns, "Impact of Winding Layer Number and
Magnet Type on Synchronous Surface PM Machines Designed for Wide
Constant-Power Speed Range Operation," Energy Conversion, IEEE
Transactions on, vol. 23, pp. 53-60, 2008.
[19] D. Ishak, Z. Q. Zhu, and D. Howe, "Permanent-magnet brushless machines with
unequal tooth widths and similar slot and pole numbers," Industry Applications,
IEEE Transactions on, vol. 41, pp. 584-590, 2005.
[20] N. Bianchi, S. Bolognani, M. D. Pre, and G. A. G. G. Grezzani, "Design
considerations for fractional-slot winding configurations of synchronous
machines," Industry Applications, IEEE Transactions on, vol. 42, pp. 997-1006,
2006.
[21] F. Magnussen and H. Lendenmann, "Parasitic effects in PM machines with
concentrated windings," in Industry Applications Conference, Fourtieth IAS
Annual Meeting. , 2005, pp. 1044-1049 Vol. 2.
[22] J . Wang, Z. P. Xia, D. Howe, and S. A. Long, "Comparative Study of 3-Phase
Permanent Magnet Brushless Machines with Concentrated, Distributed and
Modular Windings," in The 3rd IET International Conference on Power
Electronics, Machines and Drives, 2006, pp. 489-493.
[23] R. Wrobel, P. H. Mellor, N. McNeill, and D. A. Staton, "Thermal Performance
of an Open-Slot Modular-Wound Machine With External Rotor," Energy
Conversion, IEEE Transactions on, vol. 25, pp. 403-411, 2010.
[24] C. R. Steen, "Direct axis aiding permanent magnets for a laminated synchronous
motor rotor," US Patent 4139790, 1979.
[25] V. B. Honsinger, "The Fields and Parameters of Interior Type AC Permanent
Magnet Machines," Power Apparatus and Systems, IEEE Transactions on, vol.
PAS-101, pp. 867-876, 1982.
[26] M. Rahman, T. Little, and G. Slemon, "Analytical models for interior-type
permanent magnet synchronous motors," Magnetics, IEEE Transactions on, vol.
21, pp. 1741-1743, 1985.
[27] A. Consoli and G. Renna, "Interior type permanent magnet synchronous motor
analysis by equivalent circuits," Energy Conversion, IEEE Transactions on, vol.
4, pp. 681-689, 1989.
[28] B. J . Chalmers, S. A. Hamed, and G. D. Baines, "Parameters and performance of
a high-field permanent-magnet synchronous motor for variable-frequency
operation," Electric Power Applications, IEE Proceedings B, vol. 132, pp. 117-
124, 1985.
[29] R. Schifer and T. A. Lipo, "Core loss in buried magnet permanent magnet
synchronous motors," Energy Conversion, IEEE Transactions on, vol. 4, pp.
279-284, 1989.
[30] T. M. J ahns, G. B. Kliman, and T. W. Neumann, "Interior Permanent-Magnet
Synchronous Motors for Adjustable-Speed Drives," Industry Applications, IEEE
Transactions on, vol. IA-22, pp. 738-747, 1986.
References
229
[31] T. M. J ahns, "Flux-Weakening Regime Operation of an Interior Permanent-
Magnet Synchronous Motor Drive," Industry Applications, IEEE Transactions
on, vol. IA-23, pp. 681-689, 1987.
[32] W. L. Soong and T. J . E. Miller, "Field-weakening performance of brushless
synchronous AC motor drives," Electric Power Applications, IEE Proceedings -,
vol. 141, pp. 331-340, 1994.
[33] W. L. Soong, "Design and Modelling of Axially-Laminated Interior Permanent
Magnet Motor Drives for Field Weakening Applications," PhD, Deparmanent of
Electronics and Electrical Engineering, The University of Glasgow, 1993.
[34] W. L. Soong, D. A. Staton, and T. J . E. Miller, "Design of a new axially-
laminated interior permanent magnet motor," Industry Applications, IEEE
Transactions on, vol. 31, pp. 358-367, 1995.
[35] Y. Honda, T. Higaki, S. Morimoto, and Y. Takeda, "Rotor design optimisation
of a multi-layer interior permanent-magnet synchronous motor," Electric Power
Applications, IEE Proceedings -, vol. 145, pp. 119-124, 1998.
[36] Y. Honda, T. Nakamura, T. Higaki, and Y. A. Takeda, "Motor design
considerations and test results of an interior permanent magnet synchronous
motor for electric vehicles," in Industry Applications Conference, Thirty-Second
IAS Annual Meeting, IAS '97., Conference Record of the IEEE, 1997, pp. 75-82
vol.1.
[37] L. J olly, M. A. J abbar, and Q. Liu, "Optimization of the constant power speed
range of a saturated permanent-magnet synchronous motor," Industry
Applications, IEEE Transactions on, vol. 42, pp. 1024-1030, 2006.
[38] A. Fratta, A. Vagati, and F. Villata, "On the evolution of AC machines for
spindle drive applications," Industry Applications, IEEE Transactions on, vol.
28, pp. 1081-1086, 1992.
[39] S. Morimoto, Y. Takeda, and T. Hirasa, "Current phase control methods for
permanent magnet synchronous motors," Power Electronics, IEEE Transactions
on, vol. 5, pp. 133-139, 1990.
[40] Z. Q. Zhu, Y. S. Chen, and D. Howe, "Iron loss in permanent-magnet brushless
AC machines under maximum torque per ampere and flux weakening control,"
Magnetics, IEEE Transactions on, vol. 38, pp. 3285-3287, 2002.
[41] K. Kyung-Tae, K. Kwang-Suk, H. Sang-Moon, K. Tae-J ong, and J . Yoong-Ho,
"Comparison of magnetic forces for IPM and SPM motor with rotor
eccentricity," Magnetics, IEEE Transactions on, vol. 37, pp. 3448-3451, 2001.
[42] L. J ung Ho, K. J ung Chul, and H. Dong Seek, "Effect analysis of magnet on Ld
and Lq inductance of permanent magnet assisted synchronous reluctance motor
using finite element method," Magnetics, IEEE Transactions on, vol. 35, pp.
1199-1202, 1999.
[43] S. Morimoto, M. Sanada, and Y. Takeda, "Wide-speed operation of interior
permanent magnet synchronous motors with high-performance current
regulator," Industry Applications, IEEE Transactions on, vol. 30, pp. 920-926,
1994.
[44] M. N. Uddin, T. S. Radwan, G. H. George, and M. A. Rahman, "Performance of
current controllers for VSI-fed IPMSM drive," Industry Applications, IEEE
Transactions on, vol. 36, pp. 1531-1538, 2000.
[45] M. F. Rahman, L. Zhong, and L. Khiang Wee, "A direct torque-controlled
interior permanent magnet synchronous motor drive incorporating field
References
230
weakening," Industry Applications, IEEE Transactions on, vol. 34, pp. 1246-
1253, 1998.
[46] S. Ogasawara and H. Akagi, "Implementation and position control performance
of a position-sensorless IPM motor drive system based on magnetic saliency,"
Industry Applications, IEEE Transactions on, vol. 34, pp. 806-812, 1998.
[47] S. Ogasawara and H. Akagi, "An approach to real-time position estimation at
zero and low speed for a PM motor based on saliency," Industry Applications,
IEEE Transactions on, vol. 34, pp. 163-168, 1998.
[48] M. J . Corley and R. D. Lorenz, "Rotor position and velocity estimation for a
salient-pole permanent magnet synchronous machine at standstill and high
speeds," Industry Applications, IEEE Transactions on, vol. 34, pp. 784-789,
1998.
[49] M. E. Haque, Z. Limin, and M. F. Rahman, "A sensorless initial rotor position
estimation scheme for a direct torque controlled interior permanent magnet
synchronous motor drive," Power Electronics, IEEE Transactions on, vol. 18,
pp. 1376-1383, 2003.
[50] P. Sergeant, F. De Belie, and J . Melkebeek, "Effect of Rotor Geometry and
Magnetic Saturation in Sensorless Control of PM Synchronous Machines,"
Magnetics, IEEE Transactions on, vol. 45, pp. 1756-1759, 2009.
[51] K. Gyu-Hong, H. J ung-Pyo, K. Gyu-Tak, and P. J ung-Woo, "Improved
parameter modeling of interior permanent magnet synchronous motor based on
finite element analysis," Magnetics, IEEE Transactions on, vol. 36, pp. 1867-
1870, 2000.
[52] N. Bianchi and S. Bolognani, "Influence of Rotor Geometry of an IPM Motor on
Sensorless Control Feasibility," Industry Applications, IEEE Transactions on,
vol. 43, pp. 87-96, 2007.
[53] N. Bianchi, S. Bolognani, J . J i-Hoon, and S. Seung-Ki, "Comparison of PM
Motor Structures and Sensorless Control Techniques for Zero-Speed Rotor
Position Detection," Power Electronics, IEEE Transactions on, vol. 22, pp.
2466-2475, 2007.
[54] N. Bianchi, S. Bolognani, and M. Zigliotto, "Design Hints of an IPM
Synchronous Motor for an Effective Position Sensorless Control," in IEEE 36th
Power Electronics Specialists Conference, PESC, 2005, pp. 1560-1566.
[55] W. Shanshan, D. D. Reigosa, Y. Shibukawa, M. A. Leetmaa, R. D. Lorenz, and
L. Yongdong, "Interior Permanent-Magnet Synchronous Motor Design for
Improving Self-Sensing Performance at Very Low Speed," Industry
Applications, IEEE Transactions on, vol. 45, pp. 1939-1946, 2009.
[56] E. R. Braga Filho, A. M. N. Lima, and T. S. Araujo, "Reducing cogging torque
in interior permanent magnet machines without skewing," Magnetics, IEEE
Transactions on, vol. 34, pp. 3652-3655, 1998.
[57] Y. Fujishima, S. Wakao, M. Kondo, and N. Terauchi, "An optimal design of
interior permanent magnet synchronous motor for the next generation commuter
train," Applied Superconductivity, IEEE Transactions on, vol. 14, pp. 1902-1905,
2004.
[58] Y. Kano and N. Matsui, "A Design Approach for Direct-Drive Permanent-
Magnet Motors," Industry Applications, IEEE Transactions on, vol. 44, pp. 543-
554, 2008.
References
231
[59] K. Yamazaki, "Torque and efficiency calculation of an interior permanent
magnet motor considering harmonic iron losses of both the stator and rotor,"
Magnetics, IEEE Transactions on, vol. 39, pp. 1460-1463, 2003.
[60] S. Dong-J oon, C. Dong-Hyeok, C. J ang-Sung, J . Hyun-Kyo, and C. Tae-Kyoung,
"Efficiency optimization of interior permanent magnet synchronous motor using
genetic algorithms," Magnetics, IEEE Transactions on, vol. 33, pp. 1880-1883,
1997.
[61] D.-H. Kim, I.-H. Park, J .-H. Lee, and C.-E. Kim, "Optimal Shape Design of Iron
Core to Reduce Cogging Torque of IPM Motor," Magnetics, IEEE Transactions
on, vol. 39, pp. 1456-1459, 2003.
[62] K. Ki-Chan, K. Dae-Hyun, H. J ung-Pyo, and L. J u, "A Study on the
Characteristics Due to Pole-Arc to Pole-Pitch Ratio and Saliency to Improve
Torque Performance of IPMSM," Magnetics, IEEE Transactions on, vol. 43, pp.
2516-2518, 2007.
[63] L. Parsa and H. Lei, "Interior Permanent Magnet Motors With Reduced Torque
Pulsation," Industrial Electronics, IEEE Transactions on, vol. 55, pp. 602-609,
2008.
[64] A. Kioumarsi, M. Moallem, and B. Fahimi, "Mitigation of Torque Ripple in
Interior Permanent Magnet Motors by Optimal Shape Design," Magnetics, IEEE
Transactions on, vol. 42, pp. 3706-3711, 2006.
[65] K. Gyu-Hong, S. Young-Dae, K. Gyu-Tak, and H. J in, "A Novel Cogging
Torque Reduction Method for Interior-Type Permanent-Magnet Motor,"
Industry Applications, IEEE Transactions on, vol. 45, pp. 161-167, 2009.
[66] S.-H. Han, T. M. J ahns, W. L. Soong, M. K. Guven, and M. S. Illindala, "Torque
Ripple Reduction in Interior Permanent Magnet Synchronous Machines Using
Stators With Odd Number of Slots Per Pole Pair," Energy Conversion, IEEE
Transactions on, vol. 25, pp. 118-127.
[67] M. Sanada, K. Hiramoto, S. Morimoto, and Y. Takeda, "Torque ripple
improvement for synchronous reluctance motor using an asymmetric flux barrier
arrangement," Industry Applications, IEEE Transactions on, vol. 40, pp. 1076-
1082, 2004.
[68] F. Liang, K. Sung-Il, O. K. Soon, and H. J ung-Pyo, "Novel Double-Barrier
Rotor Designs in Interior-PM<newline/>Motor for Reducing Torque Pulsation,"
Magnetics, IEEE Transactions on, vol. 46, pp. 2183-2186.
[69] S.-I. Kim, J .-Y. Lee, Y.-K. Kim, J .-P. Hong, Y. Hur, and Y.-H. J ung,
"Optimization for reduction of torque ripple in interior permanent magnet motor
by using the Taguchi method," Magnetics, IEEE Transactions on, vol. 41, pp.
1796-1799, 2005.
[70] Y. Kawase, T. Ota, and H. Fukunaga, "3-D eddy current analysis in permanent
magnet of interior permanent magnet motors," Magnetics, IEEE Transactions on,
vol. 36, pp. 1863-1866, 2000.
[71] V. Zivotic-Kukolj, W. L. Soong, and N. Ertugrul, "Iron Loss Reduction in an
Interior PM Automotive Alternator," Industry Applications, IEEE Transactions
on, vol. 42, pp. 1478-1486, 2006.
[72] D. M. Ionel, M. Popescu, M. I. McGilp, T. J . E. Miller, S. J . Dellinger, and R. J .
Heideman, "Computation of Core Losses in Electrical Machines Using
Improved Models for Laminated Steel," Industry Applications, IEEE
Transactions on, vol. 43, pp. 1554-1564, 2007.
References
232
[73] K. Yamazaki and Y. Seto, "Iron loss analysis of interior permanent-magnet
synchronous motors-variation of main loss factors due to driving condition,"
Industry Applications, IEEE Transactions on, vol. 42, pp. 1045-1052, 2006.
[74] A. Wang, H. Li, and C.-T. Liu, "On the Material and Temperature Impacts of
Interior Permanent Magnet Machine for Electric Vehicle Applications,"
Magnetics, IEEE Transactions on, vol. 44, pp. 4329-4332, 2008.
[75] K. Yamazaki and Y. Kanou, "Rotor Loss Analysis of Interior Permanent Magnet
Motors Using Combination of 2-D and 3-D Finite Element Method," Magnetics,
IEEE Transactions on, vol. 45, pp. 1772-1775, 2009.
[76] T. Okitsu, D. Matsuhashi, and K. Muramatsu, "Method for Evaluating the Eddy
Current Loss of a Permanent Magnet in a PM Motor Driven by an Inverter
Power Supply Using Coupled 2-D and 3-D Finite Element Analyses," Magnetics,
IEEE Transactions on, vol. 45, pp. 4574-4577, 2009.
[77] J .-H. Seo, S.-Y. Kwak, S.-Y. J ung, C.-G. Lee, T.-K. Chung, and H.-K. J ung, "A
Research on Iron Loss of IPMSM With a Fractional Number of Slot Per Pole,"
Magnetics, IEEE Transactions on, vol. 45, pp. 1824-1827, 2009.
[78] K. Yamazaki and H. Ishigami, "Rotor-Shape Optimization of Interior-
Permanent-Magnet Motors to Reduce Harmonic Iron Losses," Industrial
Electronics, IEEE Transactions on, vol. 57, pp. 61-69.
[79] K. Yamazaki and A. Abe, "Loss Investigation of Interior Permanent-Magnet
Motors Considering Carrier Harmonics and Magnet Eddy Currents," Industry
Applications, IEEE Transactions on, vol. 45, pp. 659-665, 2009.
[80] H. Seok-Hee, W. L. Soong, T. M. J ahns, M. K. Guven, and M. S. Illindala,
"Reducing Harmonic Eddy-Current Losses in the Stator Teeth of Interior
Permanent Magnet Synchronous Machines During Flux Weakening," Energy
Conversion, IEEE Transactions on, vol. 25, pp. 441-449.
[81] K. J . Tseng and S. B. Wee, "Analysis of flux distribution and core losses in
interior permanent magnet motor," Energy Conversion, IEEE Transactions on,
vol. 14, pp. 969-975, 1999.
[82] B. Stumberger, A. Hamler, and B. Hribernik, "Analysis of iron loss in interior
permanent magnet synchronous motor over a wide-speed range of constant
output power operation," Magnetics, IEEE Transactions on, vol. 36, pp. 1846-
1849, 2000.
[83] L. Ma, M. Sanada, S. Morimoto, and Y. Takeda, "Iron loss prediction
considering the rotational field and flux density harmonics in IPMSM and
SynRM," Electric Power Applications, IEE Proceedings -, vol. 150, pp. 747-751,
2003.
[84] M. Barcaro, N. Bianchi, and F. Magnussen, "Rotor Flux-Barrier Geometry
Design to Reduce Stator Iron Losses in Synchronous IPM Motors Under FW
Operations," Industry Applications, IEEE Transactions on, vol. 46, pp. 1950-
1958, 2010.
[85] F. Fernandez-Bernal, A. Garcia-Cerrada, and R. Faure, "Determination of
parameters in interior permanent-magnet synchronous motors with iron losses
without torque measurement," Industry Applications, IEEE Transactions on, vol.
37, pp. 1265-1272, 2001.
[86] J .-Y. Lee, S.-H. Lee, G.-H. Lee, J .-P. Hong, and J . Hur, "Determination of
parameters considering magnetic nonlinearity in an interior permanent magnet
References
233
synchronous motor," Magnetics, IEEE Transactions on, vol. 42, pp. 1303-1306,
2006.
[87] Z. Li, S. Z. J iang, Z. Q. Zhu, and C. C. Chan, "Analytical Modeling of Open-
Circuit Air-Gap Field Distributions in Multisegment and Multilayer Interior
Permanent-Magnet Machines," Magnetics, IEEE Transactions on, vol. 45, pp.
3121-3130, 2009.
[88] E. C. Lovelace, T. M. J ahns, T. A. Keim, and J . H. Lang, "Mechanical design
considerations for conventionally laminated, high-speed, interior PM
synchronous machine rotors," Industry Applications, IEEE Transactions on, vol.
40, pp. 806-812, 2004.
[89] G.-B. Hwang, S.-M. Hwang, W.-B. J ung, B.-S. Kang, I. C. Hwang, and C.-U.
Kim, "Analysis of magnetic forces in magnetically saturated permanent magnet
motors by considering mechanical and magnetic coupling effects," Journal of
Applied Physics, vol. 91, pp. 6976-6978, 2002.
[90] A. M. El-Refaie, N. C. Harris, T. M. J ahns, and K. M. Rahman, "Thermal
analysis of multibarrier interior PM synchronous Machine using lumped
parameter model," Energy Conversion, IEEE Transactions on, vol. 19, pp. 303-
309, 2004.
[91] B. A. Welchko, T. M. J ahns, and S. Hiti, "IPM synchronous machine drive
response to a single-phase open circuit fault," Power Electronics, IEEE
Transactions on, vol. 17, pp. 764-771, 2002.
[92] N. Bianchi, S. Bolognani, M. Zigliotto, and M. Zordan, "Innovative remedial
strategies for inverter faults in IPM synchronous motor drives," Energy
Conversion, IEEE Transactions on, vol. 18, pp. 306-314, 2003.
[93] J . Cros, P. Viarouge, and C. Gelinas, "Design of PM brushless motors using
iron-resin composites for automotive applications," in Industry Applications
Conference, 1998. Thirty-Third IAS Annual Meeting. The 1998 IEEE, 1998, pp.
5-11 vol.1.
[94] M. Nakano and H. Kometani, "A study on eddy-current losses in rotors of
surface permanent magnet synchronous machines," in Industry Applications
Conference, 2004. 39th IAS Annual Meeting. Conference Record of the 2004
IEEE, 2004, pp. 1696-1702 vol.3.
[95] P. B. Reddy, T. M. J ahns, and A. M. El-Refaie, "Impact of Winding Layer
Number and Slot/Pole Combination on AC Armature Losses of Synchronous
Surface PM Machines Designed for Wide Constant-Power Speed Range
Operation," in Industry Applications Society Annual Meeting, 2008. IAS '08.
IEEE, 2008, pp. 1-8.
[96] H. J ussila, P. Salminen, and J . Pyrhonen, "Losses of a Permanent Magnet
Synchronous Motor with Concentrated Windings," in Power Electronics,
Machines and Drives, 2006. The 3rd IET International Conference on, 2006, pp.
207-211.
[97] H. J ussila, P. Salminen, M. Niemela, and J . Pyrhonen, "Guidelines for Designing
Concentrated Winding Fractional Slot Permanent Magnet Machines," in Power
Engineering, Energy and Electrical Drives, 2007. POWERENG 2007.
International Conference on, 2007, pp. 191-194.
[98] A. M. El-Refaie and T. M. J ahns, "Scalability of surface PM Machines with
concentrated windings designed to achieve wide speed ranges of constant-power
References
234
operation," Energy Conversion, IEEE Transaction on, vol. 21, pp. 362-369,
2006.
[99] Y. Xu and Y. Sun, "Fractional-slot low speed large torque permanent magnet
brushless motors," in 4th IEEE Conference on Industrial Electronics and
Applications, ICIEA . , 2009, pp. 3565-3569.
[100] P. Salminen, T. J okinen, and J . Pyrhonen, "Pull-out torque of fractional-slot PM-
motors with concentrated winding," Electric Power Applications, IEE
Proceedings -, vol. 152, pp. 1440-1444, 2005.
[101] P. Salminen, M. Niemela, J . Pyrhonen, and J . A. M. J . Mantere, "High-torque
low-torque-ripple fractional-slot PM-motors," in Electric Machines and Drives,
2005 IEEE International Conference on, 2005, pp. 144-148.
[102] D. Gerada, A. Mebarki, and C. Gerada, "Optimal design of a high speed
concentrated wound PMSM," in International Conference on Electrical
Machines and Systems, ICEMS, 2009, pp. 1-6.
[103] H. Murakami, H. Kataoka, Y. Honda, S. Morimoto, and Y. Takeda, "Highly
efficient brushless motor design for an air-conditioner of the next generation 42
V vehicle," in Industry Applications Conference, 2001. Thirty-Sixth IAS Annual
Meeting. Conference Record of the 2001 IEEE, 2001, pp. 461-466 vol.1.
[104] A. G. J ack, B. C. Mecrow, P. G. Dickinson, D. Stephenson, J . S. Burdess, N.
Fawcett, and J . T. Evans, "Permanent-magnet machines with powdered iron
cores and prepressed windings," Industry Applications, IEEE Transactions on,
vol. 36, pp. 1077-1084, 2000.
[105] H. Akita, Y. Nakahara, N. Miyake, and T. Oikawa, "New core structure and
manufacturing method for high efficiency of permanent magnet motors," in
Industry Applications Conference, 2003. 38th IAS Annual Meeting. Conference
Record of the, 2003, pp. 367-372 vol.1.
[106] H. Akita, Y. Nakahara, N. Miyake, and T. Oikawa, "A new core," IEEE Industry
Applications Magazine, vol. 11, pp. 38-43, 2005.
[107] W. Wang, M. E. Carrier, P. G. Michaels, J . S. Rose, and J . J . J urkowski, "End
cap for interconnecting winding coils of a segmented stator to reduce phase-on-
phase conditions and associated methods," US Patent 7116023, 2006.
[108] S. Wainio, J . Ahlvik, M. Poyhonen, and I. T. F. I. Ikonen, "Segmented stator and
the winding of it," US Patent 20040150287, 2004.
[109] K. A. Sheeran, P. Rassoolkhani, and P. G. Michaels, "Segmented Stator with
Improved Handling and Winding Characteristics," US Patent 20080129142,
2008.
[110] Y. Mitsui and K. Fujita, "Method of Manufacturing Laminated Core," US Patent
20100192357, 2010.
[111] G. E. Horst and K. I. Hoemann, "Reduced coil segmented stator," US Patent
7122933, 2006.
[112] Y. Asano, Y. Honda, H. Murakami, Y. Takeda, and S. Morimoto, "Novel noise
improvement technique for a PMSM with concentrated winding," in Power
Conversion Conference, 2002. PCC Osaka 2002. Proceedings of the, 2002, pp.
460-465 vol.2.
[113] F. Magnussen, P. Thelin, and C. Sadarangani, "Performance evaluation of
permanent magnet synchronous machines with concentrated and distributed
windings including the effect of field-weakening," in Power Electronics,
References
235
Machines and Drives, 2004. (PEMD 2004). Second International Conference on
(Conf. Publ. No. 498), 2004, pp. 679-685 Vol.2.
[114] S. O. Kwon, K. Sung-Il, Z. Peng, and H. J ung-Pyo, "Performance comparison of
IPMSM with distributed and concentrated windings," in Industry Applications
Conference, 41st IAS Annual Meeting. Conference Record of the IEEE, 2006, pp.
1984-1988.
[115] D. Ishak, Z. Q. Zhu, and D. Howe, "Comparison of PM brushless motors, having
either all teeth or alternate teeth wound," Energy Conversion, IEEE
Transactions on, vol. 21, pp. 95-103, 2006.
[116] A. M. El-Refaie, Z. Q. Zhu, T. M. J ahns, and D. Howe, "Winding Inductances of
Fractional Slot Surface-Mounted Permanent Magnet Brushless Machines," in
Industry Applications Society Annual Meeting, 2008. IAS '08. IEEE, 2008, pp. 1-
8.
[117] M. T. Abolhassani and H. A. Toliyat, "Fault tolerant permanent magnet motor
drives for electric vehicles," in Electric Machines and Drives Conference, 2009.
IEMDC '09. IEEE International, 2009, pp. 1146-1152.
[118] M. R. Shah, A. M. El-Refaie, and K. Sivasubramaniam, "Analysis of turn-to-
turn faults in surface PM machines with multi-layer fractional-slot concentrated
windings," in Electrical Machines, 2008. ICEM 2008. 18th International
Conference on, 2008, pp. 1-4.
[119] P. M. Lindh, H. K. J ussila, M. Niemela, A. Parviainen, and J . Pyrhonen,
"Comparison of Concentrated Winding Permanent Magnet Motors With
Embedded and Surface-Mounted Rotor Magnets," Magnetics, IEEE
Transactions on, vol. 45, pp. 2085-2089, 2009.
[120] P. Salminen, J . Pyrhonen, H. J ussila, and M. Niemela, "Concentrated Wound
Permanent Magnet Machines with Different Rotor Designs," in Power
Engineering, Energy and Electrical Drives, 2007. POWERENG 2007.
International Conference on, 2007, pp. 514-517.
[121] S. H. Han, T. M. J ahns, and Z. Q. Zhu, "Analysis of Rotor Core Eddy-Current
Losses in Interior Permanent Magnet Synchronous Machines," in Industry
Applications Society Annual Meeting, IAS '08. IEEE, 2008, pp. 1-8.
[122] Y. Kawaguchi, T. Sato, I. Miki, and M. Nakamura, "A reduction method of
cogging torque for IPMSM," in Electrical Machines and Systems, 2005. ICEMS
2005. Proceedings of the Eighth International Conference on, 2005, pp. 248-250
Vol. 1.
[123] Y. Kano, T. Terahai, T. Kosaka, N. Matsui, and T. Nakanishi, "A new flux-
barrier design of torque ripple reduction in saliency-based sensorless drive IPM
motors for general industrial applications," in Energy Conversion Congress and
Exposition, 2009. ECCE 2009. IEEE, 2009, pp. 1939-1945.
[124] Y. Kano, T. Kosaka, N. Matsui, and T. Nakanishi, "A new technique of torque
ripple reduction in saliency-based sensorless drive IPM motors for general
industrial applications," in 13th European Conference on Power Electronics and
Applications, EPE '09. , 2009, pp. 1-10.
[125] S.-I. Kim, J .-H. Bhan, J .-P. Hong, and K.-C. Lim, "Optimization Technique for
Improving Torque Performance of Concentrated Winding Interior PM
Synchronous Motor with Wide Speed Range," in Industry Applications
Conference, 41st IAS Annual Meeting. Conference Record of the 2006 IEEE,
2006, pp. 1933-1940.
References
236
[126] K.-j. Lee, K. Kim, S. Kim, J .-S. Ahn, S. Lim, and J . Lee, "Optimal magnet shape
to improve torque characteristics of interior permanent magnet synchronous
motor," Journal of Applied Physics, vol. 97, pp. 10Q505-10Q505-3, 2005.
[127] S. Y. Lim and J . Lee, "A design for improved performance of interior permanent
magnet synchronous motor for hybrid electric vehicle," Journal of Applied
Physics, vol. 99, pp. 08R308-08R308-3, 2006.
[128] K.-C. Kim, K. Kim, H. J . Kim, and J . Lee, "Demagnetization Analysis of
Permanent Magnets According to Rotor Types of Interior Permanent Magnet
Synchronous Motor," Magnetics, IEEE Transactions on, vol. 45, pp. 2799-2802,
2009.
[129] A. M. El-Refaie and T. M. J ahns, "Comparison of synchronous PM machine
types for wide constant-power speed range operation," in Industry Applications
Conference, Fourtieth IAS Annual Meeting, 2005, pp. 1015-1022 Vol. 2.
[130] A. M. El-Refaie and T. M. J ahns, "Optimal flux weakening in surface PM
machines using fractional-slot concentrated windings," Industry Applications,
IEEE Transactions on, vol. 41, pp. 790-800, 2005.
[131] A. M. El-Refaie, T. M. J ahns, P. J . McCleer, and J . W. McKeever,
"Experimental verification of optimal flux weakening in surface PM Machines
using concentrated windings," Industry Applications, IEEE Transactions on, vol.
42, pp. 443-453, 2006.
[132] W. L. Soong, P. B. Reddy, A. M. El-Refaie, T. M. J ahns, and N. Ertugrul,
"Surface PM Machine Parameter Selection for Wide Field-Weakening
Applications," in Industry Applications Conference, 2007. 42nd IAS Annual
Meeting. Conference Record of the 2007 IEEE, 2007, pp. 882-889.
[133] A. R. Munoz, F. Liang, and M. W. Degner, "Evaluation of Interior PM and
Surface PM Synchronous machines with distributed and concentrated windings,"
in 34th Annual Conference of IEEE Industrial Electronics, IECON. , 2008, pp.
1189-1193.
[134] C. Deak, A. Binder, B. Funieru, and M. Mirzaei, "Extended field weakening and
overloading of high-torque density permanent magnet motors," in Energy
Conversion Congress and Exposition, 2009. ECCE 2009. IEEE, 2009, pp. 2347-
2353.
[135] P. H. Mellor, R. Wrobel, and D. Holliday, "A computationally efficient iron loss
model for brushless AC machines that caters for rated flux and field weakened
operation," in IEEE International Electric Machines and Drives Conference,
IEMDC '09. , 2009, pp. 490-494.
[136] F. Meier and J . Soulard, "Analysis of Flux Measurements on a PMSM With
Non-Overlapping Concentrated Windings," in IEEE Industry Applications
Society Annual Meeting, IAS '08., 2008, pp. 1-8.
[137] R. Nuscheler, "Two-dimensional analytical model for eddy-current loss
calculation in the magnets and solid rotor yokes of permanent magnet
synchronous machines," in 18th International Conference on Electrical
Machines, ICEM. , 2008, pp. 1-6.
[138] O. Bottauscio, G. Pellegrino, P. Guglielmi, M. Chiampi, and A. Vagati, "Rotor
loss estimation in permanent magnet machines with concentrated windings,"
Magnetics, IEEE Transactions on, vol. 41, pp. 3913-3915, 2005.
References
237
[139] M. Nakano, H. Kometani, and M. Kawamura, "A study on eddy-current losses
in rotors of surface permanent-magnet synchronous machines," Industry
Applications, IEEE Transactions on, vol. 42, pp. 429-435, 2006.
[140] K. Yamazaki, M. Shina, Y. Kanou, M. Miwa, and J . Hagiwara, "Effect of Eddy
Current Loss Reduction by Segmentation of Magnets in Synchronous Motors:
Difference Between Interior and Surface Types," Magnetics, IEEE Transactions
on, vol. 45, pp. 4756-4759, 2009.
[141] K. Yamazaki, Y. Fukushima, and M. Sato, "Loss Analysis of Permanent Magnet
Motors With Concentrated Windings - Variation of Magnet Eddy Current Loss
Due to Stator and Rotor Shapes," Industry Applications, IEEE Transactions on,
vol. 45, pp. 1334- 1342 2009.
[142] H. Polinder, M. J . Hoeijmakers, and M. Scuotto, "Eddy-current losses in the
solid back-iron of permanent magnet machines with concentrated fractional
pitch windings," in The 3rd IET International Conference on Power Electronics
Machines and Drives. PEMD 2006. , 2006, pp. 479-483.
[143] J . Wang, Z. P. Xia, D. Howe, and S. A. Long, "Vibration Characteristics of
Modular Permanent Magnet Brushless AC Machines," in IEEE Industry
Applications Conference, 41st IAS Annual Meeting. , 2006, pp. 1501-1506.
[144] S.-H. Lee and J .-P. Hong, "A Study on the Acoustic Noise Reduction of Interior
Permanent Magnet Motor with Concentrated Winding," in Industry Applications
Society Annual Meeting, 2008. IAS '08. IEEE, 2008, pp. 1-5.
[145] S.-H. Lee, J .-P. Hong, S.-M. Hwang, W.-T. Lee, J .-Y. Lee, and Y.-K. Kim,
"Optimal Design for Noise Reduction in Interior Permanent-Magnet Motor,"
Industry Applications, IEEE Transactions on, vol. 45, pp. 1954-1960, 2009.
[146] S. Araki, T. Kondou, and A. Yamagiwa, "Analyses of electromagnetic forces of
concentrated winding permanent magnet brushless motors with rotor
eccentricity," in International Conference on Electrical Machines and Systems,
ICEMS, 2009, pp. 1-4.
[147] F. Xu and T. Li, "A research on the radial unbalanced force in the
unsymmetrical brushless PM motors," in Industrial Electronics and Applications,
2008. ICIEA 2008. 3rd IEEE Conference on, 2008, pp. 1695-1698.
[148] W. J ingai, Z. Dongqi, and A. Satake, "The torque research on concentrated
winding brushless PM motor," in TENCON '02. Proceedings. 2002 IEEE Region
10 Conference on Computers, Communications, Control and Power
Engineering, 2002, pp. 2026-2029 vol.3.
[149] R. Qu and T. A. Lipo, "General closed-form analytical expressions of air-gap
inductances for surface-mounted permanent magnet and induction machines," in
Electric Machines and Drives Conference, 2003. IEMDC'03. IEEE International,
2003, pp. 443-448 vol.1.
[150] I. Abdennadher, R. Kessentini, and A. Masmoudi, "Analytical derivation and
FEA validation of the inductances of CWPMM," in 6th International Multi-
Conference on Systems, Signals and Devices, SSD, 2009, pp. 1-7.
[151] J . K. Tangudu, T. M. J ahns, A. El-Refaie, and Z. Q. Zhu, "Lumped parameter
magnetic circuit model for fractional-slot concentrated-winding interior
permanent magnet machines," in Energy Conversion Congress and Exposition,
2009. ECCE 2009. IEEE, 2009, pp. 2423-2430.
[152] J . K. Tangudu, T. M. J ahns, A. M. El-Refaie, and Z. Q. Zhu, "Segregation of
torque components in fractional-slot concentrated-winding interior PM machines
References
238
using frozen permeability," in Energy Conversion Congress and Exposition,
2009. ECCE 2009. IEEE, 2009, pp. 3814-3821.
[153] F. Meier and J . Soulard, "dq theory applied to a permanent magnet synchronous
machine with concentrated windings," in 4th IET Conference on Power
Electronics, Machines and Drives. PEMD. , 2008, pp. 194-198.
[154] Y. Duan, R. G. Harley, and T. G. Habetler, "Method for multi-objective
optimized designs of Surface Mount Permanent Magnet motors with
concentrated or distributed stator windings," in Electric Machines and Drives
Conference, 2009. IEMDC '09. IEEE International, 2009, pp. 323-328.
[155] D. Reigosa, K. Akatsu, N. Limsuwan, Y. Shibukawa, and R. D. Lorenz, "Self-
sensing comparison of fractional slot pitch winding vs. distributed winding for
FW- and FI-IPMSMs based on carrier signal injection at very low speed," in
Energy Conversion Congress and Exposition, 2009. ECCE 2009. IEEE, 2009,
pp. 3806-3813.
[156] A. Eilenberger, M. Schroedl, and J . Heissenberger, "Comparison of outer rotor
PMSM with single- and double-layer windings at same machine geometry with
respect to the sensorless control capability," in 13th European Conference on
Power Electronics and Applications, EPE, 2009, pp. 1-7.
[157] A. Eilenberger, M. Schroedl, F. Demmelmayr, and M. Troyer, "Short-circuit-
proofed outer rotor PMSM with a wide field weakening range for high
efficiency traction applications," in Industrial Electronics, 2009. IECON '09.
35th Annual Conference of IEEE, 2009, pp. 1294-1297.
[158] N. Imai, S. Morimoto, M. Sanada, and Y. Takeda, "Influence of Rotor
Configuration on Sensorless Control for Permanent-Magnet Synchronous
Motors," Industry Applications, IEEE Transactions on, vol. 44, pp. 93-100,
2008.
[159] K. Kojima, M. Hasegawa, T. Ohnuma, M. Tomita, S. Doki, S. Okuma, and K.
Matsui, "Position sensorless control by reduced-order extended-flux observer
applicable to IPMSM with concentrated windings," in Power Electronics
Electrical Drives Automation and Motion (SPEEDAM), 2010 International
Symposium on, pp. 1018-1023.
[160] Y. Wang, W. Xuhui, S. Xue, T. Fan, and Z. Lili, "Analysis and design of high
power factor interior permanent magnet motor with concentrated windings for
undersea vehicle propulsion," in IEEE Vehicle Power and Propulsion
Conference, VPPC 2008, pp. 1-6.
[161] D. M. Ionel, "High-efficiency variable-speed electric motor drive technologies
for energy savings in the US residential sector," in 12th International
Conference on Optimization of Electrical and Electronic Equipment (OPTIM)
2010, pp. 1403-1414.
[162] M. V. Cistelecan and M. Popescu, "Study of the Number of Slots/Pole
Combinations for Low Speed Permanent Magnet Synchronous Generators," in
Electric Machines & Drives Conference, 2007. IEMDC '07. IEEE International,
2007, pp. 1616-1620.
[163] E. V. Kazmin, E. A. Lomonova, and J . J . H. Paulides, "Brushless traction PM
machines using commercial drive technology, Part II: Comparative study of the
motor configurations," in International Conference on Electrical Machines and
Systems, ICEMS., 2008, pp. 3772-3780.
References
239
[164] L. Alberti, M. Barcaro, M. D. Pre, A. Faggion, L. Sgarbossa, N. Bianchi, and S.
Bolognani, "IPM Machine Drive Design and Tests for an Integrated Starter-
Alternator Application," Industry Applications, IEEE Transactions on, vol. 46,
pp. 993-1001, 2010.
[165] J . Germishuizen and M. Kamper, "Design and Performance Characteristics of
IPM Machines with Single Layer Non-Overlapping Concentrated Windings," in
Industry Applications Conference, 2007. 42nd IAS Annual Meeting. Conference
Record of the 2007 IEEE, 2007, pp. 141-147.
[166] J . J . Germishuizen and M. J . Kamper, "IPM Traction Machine With Single
Layer Non-Overlapping Concentrated Windings," Industry Applications, IEEE
Transactions on, vol. 45, pp. 1387-1394, 2009.
[167] A. M. El-Refaie, "Fractional-Slot Concentrated-Windings Synchronous
Permanent Magnet Machines: Opportunities and Challenges," Industrial
Electronics, IEEE Transactions on, vol. 57, pp. 107-121, 2010.
[168] R. D. Cook, D. S. Malkus, M. E. Plesha, and R. J . Witt, Concepts and
Applications of Finite Element Analysis, 4th ed.: New York, NY : Wiley, 2002.
[169] S. V. Ahamed and E. A. Erdelyi, "Nonlinear Theory of Salient Pole Machines,"
Power Apparatus and Systems, IEEE Transactions on, vol. PAS-85, pp. 61-70,
1966.
[170] E. A. Erdelyi, S. V. Ahamed, and R. D. Burtness, "Flux Distribution in Saturated
DC Machines at No-Load," Power Apparatus and Systems, IEEE Transactions
on, vol. 84, pp. 375-381, 1965.
[171] P. Silvester and M. V. K. Chari, "Finite Element Solution of Saturable Magnetic
Field Problems," Power Apparatus and Systems, IEEE Transactions on, vol.
PAS-89, pp. 1642-1651, 1970.
[172] M. V. K. Chari and P. Silvester, "Finite-Element Analysis of Magnetically
Saturated D-C Machines," Power Apparatus and Systems, IEEE Transactions on,
vol. PAS-90, pp. 2362-2372, 1971.
[173] J . P. A. Bastos and N. Sadowski, Electromagnetic Modelling By Finite Element
Methods. New York, NY: Marcel Dekker, Inc., 2003.
[174] R. Dutta, "A Segmented Interior Permanent Magnet Synchronous Machine with
Wide Field-weakening Range," PhD Thesis, School of Electrical Engineering
and Telecommunications, University of New South Wales, Australia, 2007.
[175] D. J ing, "Numerical Analysis of a Permanent Magnet Synchronous Machine
Using Numerical Techniques," PhD Thesis, Department of Electrical and
Computer Engineering, National University of Singapore, Singapore, 2004.
[176] D. Zhong, "Finite Element Analysis of Synchronous Machines," PhD Thesis,
Department of Electrical Engineering, The Pennsylvania State University, 2010.
[177] J . J in, The Finite Element Method in Electromagnetics. New York, NY: J ohn
Wiley & Sons, 2002.
[178] N. Bianchi, Electrical Machine Analysis Using Finite Elements. Boca Raton, FL:
Taylor and Francis Group, LLC, 2005.
[179] P. P. Silvester and R. L. Ferrari, Finite Elements for Electrical Engineers:
Cambridge University Press, 1996.
[180] M. V. K. Chari and P. Silvester, "Analysis of Turboalternator Magnetic Fields
by Finite Elements," Power Apparatus and Systems, IEEE Transactions on, vol.
PAS-90, pp. 454-464, 1971.
References
240
[181] E. P. Furlani, Permanent Magnet and Electromechanical Devices. Rochester,
NY: Academic Press, 2001.
[182] CEDRAT, "Flux 9.30 User Guide Vol. 2 - Physical description, Circuit
Coupling, Kinematic Coupling " 2006, April.
[183] D. C. Hanselman, "Minimum torque ripple, maximum efficiency excitation of
brushless permanent magnet motors," Industrial Electronics, IEEE Transactions
on, vol. 41, pp. 292-300, 1994.
[184] A. Chiba, F. Nakamura, T. Fukao, and M. A. Rahman, "Inductances of cageless
reluctance-synchronous machines having nonsinusoidal space distributions,"
Industry Applications, IEEE Transactions on, vol. 27, pp. 44-51, 1991.
[185] A. M. El-Refaie, T. M. J ahns, and D. W. Novotny, "Analysis of surface
permanent magnet machines with fractional-slot concentrated windings," Energy
Conversion, IEEE Transaction on, vol. 21, pp. 34-43, 2006.
[186] D. A. Staton, T. J . E. Miller, and S. E. Wood, "Maximising the saliency ratio of
the synchronous reluctance motor," Electric Power Applications, IEE
Proceedings., vol. 140, pp. 249-259, 1993.
[187] L. Chong and M. F. Rahman, "Saliency ratio optimization in an IPM machine
with fractional-slot concentrated windings," in International Conference on
Electrical Machines and Systems, ICEMS. , 2008, pp. 2921-2926.
[188] L. Chong and M. F. Rahman, "Saliency ratio derivation and optimisation for an
interior permanent magnet machine with concentrated windings using finite-
element analysis," Electric Power Applications, IET, vol. 4, pp. 249-258, 2010.
[189] T. J . E. Miller, Brushless permanent-magnet and reluctance motor drives. New
York: Oxford University Press, 1989.
[190] W. L. Soong, S. Han, and T. M. J ahns, "Design of interior PM machines for
field-weakening applications," in Electrical Machines and Systems, 2007.
ICEMS. International Conference on, 2007, pp. 654-664.
[191] E. P. Furlani, Permanent Magnet and Electromechanical Devices - Materials,
Analysis and Applications: Academic Press, 2001.
[192] A. Goldman, Modern Ferrite Technology, Second ed.: Springer, 2006.
[193] (24/03/2001). Magnetic Component Engineering: Neodymium Iron Boron vs.
Samarium Cobalt- A Comparison. Available:
http://www.mceproducts.com/knowledge-base/article/article-dtl.asp?id=32
[194] "Magnet Sales and Manufacturing Inc. - High Performance Permanent Magnets
7," ed. California, 1995.
[195] Neodymium Magnet Physical Properties, K&J Magnetics Inc. [Online].
Available: http://www.kjmagnetics.com/specs.asp
[196] Magnetic Properties of Sintered NdFeB Magnets [Online]. Available:
http://www.raremagnet.com/en/product/product1_1.asp
[197] "Advaced MotorTech, LLC: Short Course on the Design of Interior Permanent
Magnet and Brushless DC Machines- Taking Theory to Practice," ed, May 2010.
[198] J FE, "Electrical Steel Data Sheets - J FE G-CORE, J FE N-CORE," ed. J apan,
2007.
[199] E. S. Hamdi, Design of Small Electrical Machines. UK: J ohn Wiley & Sons,
1994.
[200] DuPont. (2003, 16 Oct 2009). Nomex Type 410 data sheet.
References
241
[201] L. Chong, R. Dutta, and M. F. Rahman, "Field Weakening Performance of a
Concentrated Wound PM Machine with Rotor and Magnet Geometry Variation
" in Power Engineering Society General Meeting, PES, Minneapolis, USA, 2010.
[202] N. Bianchi, S. Bolognani, and P. Frare, "Design criteria of high efficiency SPM
synchronous motors," in Electric Machines and Drives Conference, IEMDC'03.
IEEE International, 2003, pp. 1042-1048 vol.2.
[203] A. Borisavljevic, H. Polinder, and J . A. Ferreira, "On the Speed Limits of
Permanent-Magnet Machines," Industrial Electronics, IEEE Transactions on,
vol. 57, pp. 220-227, 2010.
[204] T. Kikuchi and T. Kenjo. (1997, 22/03/2011). A Unique Desk-top Electrical
Machinery Laboratory for the Mechatronics Age. Available:
http://ewh.ieee.org/soc/es/Nov1997/09/INDEX.HTM
[205] N. Bianchi, S. Bolognani, and F. Luise, "Potentials and limits of high speed PM
motors," in Industry Applications Conference, 2003. 38th IAS Annual Meeting.
Conference Record of the, 2003, pp. 1918-1925 vol.3.
[206] C. Han-Wook, J . Seok-Myeong, and C. Sang-Kyu, "A Design Approach to
Reduce Rotor Losses in High-Speed Permanent Magnet Machine for Turbo-
Compressor," Magnetics, IEEE Transactions on, vol. 42, pp. 3521-3523, 2006.
[207] A. Beiser, Schaum's outline of theory and problems of physical science:
McGraw-Hill, 1988.
[208] K. Shinichi, T. Norio, and K. Takeharu, "Measurement and analysis of AC loss
of NdFeB sintered magnet," Electrical Engineering in Japan, vol. 154, pp. 8-15,
2006.
[209] CEDRAT, "Flux 9.30 User Guide Vol. 3 - Physical Applications (Magnetic,
Electric, Thermal)," 2006, April.
[210] "AK Steel - Nonoriented Electrical Steels Datasheet (M-15 through M-47)," ed,
2007.
[211] A. M. El-Refaie and T. M. J ahns, "Impact of Winding Layer Number and
Magnet Type on Synchronous Surface PM Machines Designed for Wide
Constant-Power Speed Range Operation," in Industry Applications Conference,
41st IAS Annual Meeting., 2006, pp. 1486-1493.
[212] W. H. Yeadon and A. W. Yeadon, Handbook of Small Electric Motors:
McGraw-Hill, 2001.
[213] "Sankey - 35RM300 Nonoriented Silicon Steel Datasheet," ed, 2010.
[214] H. J ussila, P. Salminen, A. Parviainen, J . Nerg, and J . Pyrhonen, "Concentrated
winding axial flux permanent magnet motor with plastic bonded magnets and
sintered segmented magnets," in 18th International Conference on Electrical
Machines, ICEM, 2008, pp. 1-5.
[215] N. Takahashi, H. Shinagawa, D. Miyagi, Y. Doi, and K. Miyata, "Factors
affecting eddy current losses of segmented Nd-Fe-B sintered magnets without
insulation in large PM motors," in IEEE International Electric Machines and
Drives Conference, IEMDC. , 2009, pp. 24-29.
[216] A. Fukuma, S. Kanazawa, D. Miyagi, and N. Takahashi, "Investigation of AC
loss of permanent magnet of SPM motor considering hysteresis and eddy-
current losses," Magnetics, IEEE Transactions on, vol. 41, pp. 1964-1967, 2005.
[217] K. Akatsu, K. Narita, Y. Sakashita, and T. Yamada, "Characteristics comparison
between SPMSM and IPMSM under high flux density condition by both
References
242
experimental and analysis results," in Electrical Machines and Systems, 2008.
ICEMS 2008. International Conference on, 2008, pp. 2848-2853.
[218] D. Ishak, Z. Q. Zhu, and D. Howe, "Eddy-current loss in the rotor magnets of
permanent-magnet brushless machines having a fractional number of slots per
pole," Magnetics, IEEE Transactions on, vol. 41, pp. 2462-2469, 2005.
[219] P. Sergeant and A. Van den Bossche, "Segmentation of Magnets to Reduce
Losses in Permanent-Magnet Synchronous Machines," Magnetics, IEEE
Transactions on, vol. 44, pp. 4409-4412, 2008.
[220] J uha Pyrhonen, Tapani J okien, and V. Hrabovcova, Design of Rotating
Electrical Machines: J ohn Wiely & Sons, Ltd., 2008.
[221] (15 May). PowerStream Technologies - American Wire Guage. Available:
http://www.powerstream.com/Wire_Size.htm
[222] Wikipedia. (2011). American Wire Guage. Available:
http://en.wikipedia.org/wiki/American_wire_gauge
[223] A. Emadi, Energy-Efficient Electric Motors, Third ed. New York: Marcel
Dekker, Inc., 2005.
[224] R. H. Park, "Two-reaction theory of synchronous machines generalized method
of analysis-part I," American Institute of Electrical Engineers, Transactions of
the, vol. 48, pp. 716-727, 1929.
[225] D. W. Novotny and T. A. Lipo, Vector control and dynamics of AC drives. New
York: Oxford University Press, 1996.
[226] M. F. Rahman, Electrical Machines and Drives Lecture Notes - Lecture 6: CSI
of Synchronous Motor Drive. Australia: University of New South Wales, 2007.
[227] P. Vas, Vector Control of AC Machines. New York: Oxford University Press,
1990.
[228] P. C. Krause, O. Wasynczuk, and S. D. Sudhoff, Analysis of Electric Machinery.
New York: IEEE Press, 1995.
[229] M. F. Rahman, Electrical Machines and Drives Lecture Notes - Lecture 9:
Dynamic Model of the Three-Phase Synchronous Motor and its Vector Control.
Australia: The University of New South Wales, 2007.
[230] S. Morimoto, Y. Takeda, T. Hirasa, and K. Taniguchi, "Expansion of operating
limits for permanent magnet motor by current vector control considering
inverter capacity," Industry Applications, IEEE Transactions on, vol. 26, pp.
866-871, 1990.
243
APPENDIX A
AC STANDSTILL TEST APPLIED TO THE FINITE ELEMENT MODEL OF THE
SEGMENTED IPMMACHINE
A.1 Results of AC Standstill Test Implemented on the Segmented IPM Machine
The AC standstill test was used for measuring inductances of the CW-IPM machine, as
it takes into account the additional leakage harmonic components. This test was
implemented in FE analysis and to determine its validity, the test carried out on the
UNSW segmented IPM machine model, as shown in fig. A.1. The results obtained from
the model was compared with other FE methods as well as the experimental AC
standstill test method. Results obtained was shown in fig. A.2.
Fig. A.1 Segmented IPM machine
Appendix A
244
Fig. A.2 AC standstill test on the S-IPM machine using FE analysis compared to other methods
I
n
d
u
c
t
a
n
c
e
I
n
d
u
c
t
a
n
c
e
245
APPENDIX B
SALIENCY RATIO OPTIMISATION
B.1 Optimisation of Saliency Ratio by Variation of Rotor Magnet Shape
Since it is difficult to shape magnet poles simply by the observation of d and q-axis
paths, there are no references and rules on how the rotor pole for a CW-IPM machine
should be shaped. Thus, randomized sets of rotor magnet shapes were modelled in order
to establish certain basic rules in maximising saliency.
In the following four designs, the magnet width and thickness were kept constant to
investigatethe effects of varying magnet and link section geometry.
D
e
s
i
g
n

1
Saliency Ratio
L
d
= 5.649 mH
L
q
= 5.831 mH
( =
L
q
L
d
= 1.032
Design Constants
Magnet width =12mm
Magnet Depth =2mm
D
e
s
i
g
n

2
Saliency Ratio
L
d
= 5.517mH
L
q
= 5.768mH
=
L
q
L
d
= 1.045
Design Constants
Fluxguide Design
Magnet width =12mm
Magnet thickness =2mm
D
e
s
i
g
n

3
Saliency Ratio
L
d
= 6.035mH
L
q
= 6.188 mH
=
L
q
L
d
= 1.025
Design Constants
v-shaped magnets
Magnet width =12mm
Magnet thickness =2mm
D
e
s
i
g
n

4
Saliency Ratio
L
d
= 6.4274 mH
L
q
= 6.4956mH
=
L
q
L
d
= 1.011
Design Constants
Segmented magnets
Magnet width =12mm
Magnet thickness =2mm
Appendix B
246
It can be observed that the variation of magnet shapes did not have much effect on
increasing the saliency ratio.
Inthe following designs (5 and 6), the tips of the flux guides were placed closer to the
rotor surface to channel more magnet flux across the airgap. This reduced the overall
inductance and increased the saliency ratio. It is also shown that an increase in magnet
thickness also caused an increase in saliency ratio.
D
e
s
i
g
n

5
Saliency Ratio
L
d
= 4.5285mH
L
q
= 4.8416mH
=
L
q
L
d
= 1.069
Design Constants
Magnet width =12mm
Magnet thickness =2mm
D
e
s
i
g
n

6
Saliency Ratio
L
d
= 4.173mH
L
q
= 4.5013mH
=
L
q
L
d
= 1.073
Design Constants
Magnet width =12mm
Magnet thickness =4mm
In the next set of designs, wider magnets were used. This resulted in an obvious
increase in saliency ratio. Various magnet shapes, positions, different number of barriers
and different magnet thickness were experimented. The outcome was that a single-
barrier structure with a basic rectangular shaped produced higher saliency ratio.
Appendix B
247
D
e
s
i
g
n

7
Saliency Ratio
L
d
= 4.1938mH
L
q
= 4.402mH
=
L
q
L
d
= 1.106
Design Constants
Magnet width =14mm
Magnet thickness =4mm
D
e
s
i
g
n

8
Saliency Ratio
L
d
= 4.1058mH
L
q
= 4.4688mH
=
L
q
L
d
= 1.088
Design Constants
Magnet width =14mm
Magnet thickness =4mm
D
e
s
i
g
n

9
Saliency Ratio
L
d
= 4.3026mH
L
q
= 4.7516mH
=
L
q
L
d
= 1.1044
Design Constants
Magnet width =14mm
Magnet thickness =4mm
D
e
s
i
g
n

1
0
Saliency Ratio
L
d
= 4.0562mH
L
q
= 4.4393mH
=
L
q
L
d
= 1.095
Design Constants
Double-barrier magnets
Magnet width =14mm
Magnet thickness =4mm
D
e
s
i
g
n

1
1
Saliency Ratio
L
d
= 4.1938mH
L
q
= 4.6402mH
=
L
q
L
d
= 1.1064
Design Constants
Magnet width =14mm
Magnet thickness =4mm
D
e
s
i
g
n

1
2
Saliency Ratio
L
d
= 3.6536mH
L
q
= 4.1544mH
=
L
q
L
d
= 1.137
Design Constants
Magnet width =14mm
Magnet thickness =6mm
Although the highest saliency achieved was just above 1.13, the various designs did
show variations in the saliency ratio. This allowed basic design rules to be highlighted
via observation of the results obtained.
Appendix B
248
Thick magnets (Design 12) help in reducing d-axis inductance, thus increasing
saliency ratio
Having magnets further away from airgap increases both d-and q-axis
inductances therefore magnets should be as close to the airgap as possible,
alternatively non-magnetic barriers at magnet edges can be extended closer
towards the airgap.
Segmentation of magnets (Design 4) increases d-axis inductance hence lowers
saliency ratio
Variation of magnet shape (Design 8 and 9) including the addition of a barrier
(Design 10) has little effect on the saliency ratio
249
APPENDIX C
INDUCTANCE WAVEFORMS AND SALIENCY RATIO FOR VARIOUS SLOT/POLE
COMBINATION AND FOR DOUBLE-LAYER WINDINGS
C.1 Inductance Waveform and Saliency Ratio Comparison of Various
Slot/pole Combinations
The saliency ratio of three different single-layer winding layouts, (with 6, 12 and 18
slots), are compared. In these three models the same 14-pole rotor is used. The AC
standstill test is used to determine the inductance waveforms of the three models.
Fig. C.1 Three 14-pole layouts used in the saliency ratio comparison
The following waveforms are for the 6, 12 and 18-slot, 14-pole, single-layer windings
as shown in the following figures:
Fig. C.2 Self- and Mutual-inductance for the 6-slot, 14-pole, single-layer model
Appendix C
250
Fig. C.3 Self- and Mutual-inductance for the12-slot, 14-pole, single-layer model
Fig. C.4 Self- and Mutual-inductance for the18-slot, 14-pole, single-layer model
Appendix C
251
C.2 Inductance Waveform and Saliency Ratio Comparison with Double-Layer
Windings
Here the saliency ratio of a 6-slot, single-layer winding machine is compared with 6-slot
double-layer winding machine. The same 14-pole rotor is used.
Fig. C.5 Single- and double-layer layout used in the saliency ratio comparison
The inductance waveform for 6-slot, single-layer model was shown in fig. C.2. The
following waveform is for the 6-slot, double-layer model:
Fig. C.6 Self- and Mutual-inductance for the 6-slot, 14-pole, double-layer model
Appendix C
252
Table C.1 Summarises the achieved saliency ratios for all the above mentioned models
Table C.1
Saliency ratios for various layouts
Layout
6-slot, 14-pole, Single-layer 1.059
12-slot, 14-pole, Single-layer 1.056
18-slot, 14-pole, Single-layer 1.087
6-slot, 14-pole, Double-layer 1.112
253
APPENDIX D
THERMAL MODEL
D.1 Thermal Model Approximating Temperature at Various Parts of the Machine
A basic thermal model was created using Motor-CAD to determine the approximate
temperature distribution in the machine. Key specifications for this thermal model is
stated in table D.1:
Table D.1
Key machine parameters used in thermal model
Stator/Rotor material N.O. Silicon Steel
Core loss @50Hz/1.5T 3.40W/kg
Winding material Copper
Total no. of turns per coils 228 Turns
Stator voltage 30V
rms
Stator resistance 0.18 O
Stator current 17A
rms
Output power 1kW
Airgap length 0.6mm
Rotor radius 42mm
Machine effective length 65mm
Machine radius 65mm
Slot fill-factor 0.5
Housing radial thickness 2.5mm
Conductor thickness AWG17
Ambient Temperature 30C
Airgap thickness 0.6mm
Cooling arrangements Naturally cooled (no fins)
Appendix D
254
Fig. D.1 Approximate model used in thermal analysis
The model as shown in fig D.1 was set to operate with 250Hz excitation frequency with
an ambient temperature of 30C. It was subjected to the worse case scenario where the
machine was completely enclosed in the case (without fins) and made to run
continuously at full load current over a timeframe of 20 mins. The steady state
temperatures at various parts of the machine are given in table D.2:
Table D.2
Estimated temperature at various parts of the machine
Housing 94C
Rotor yokeand rotor surface 107C
Magnet 107C
Stator surfaceand stator tooth 103C
Stator yoke 100C
Winding 121C
Housing 94C
Rotor yoke
Housing
Rotor surface
Magnet
Stator tooth
Winding
Stator surface
Stator yoke
255
APPENDIX E
FINAL MACHINE DRAWINGS
E.1 ABB Casing used (with Original Induction Motor)
Appendix E
256
E.2 Stator of the CW-IPM Prototype
Appendix E
257
E.3 Rotor of the CW-IPM Prototype
Appendix E
258
Appendix E
259
E.4 Shaft of the CW-IPM Prototype
Appendix E
260
E.5 Key (shaft) of the CW-IPM Prototype
Appendix E
261
E.6 End-plates of the CW-IPM Prototype
262
APPENDIX F
EXPERIMENTAL SETUP
F.1 The Experimental Setup
The experimental setup of for testing the performance of the CW-IPM prototype,
(shown in fig. F.1), consists of the following:
(i) CW-IPM machine
(ii) Kollmorgen PM machine (loading machine)
(iii) Load resistor bank (1.1O - 116.0O)
(iv) Windows based PC
(v) DS1104 Controller board
(vi) 3-phase IGBT inverter
(vii) 10Nm Torque transducer (HBM T20WN)
(viii) PositionSensor (Heidenhain ROD 426 5000 pulses per revolution)
(ix) Power analyser
(x) 415V Voltage regulator
Fig. F.1 Complete experimental Setup
(i)
(viii)
(vi)
(x)
(vi)
(iii)
(ix)
(ii)
(vii)
Appendix F
263
In this setup, the inverter was supplied by 3-phase, 415V mains through the voltage
regulator. The DS1104 controller board receives feedback signals from the inverter,
(supply current from two phases), as well as from the position sensor. These signals,
together with the applied control algorithm and SVM produces the desired current
references to the inverter, (through the controller board). The controller algorithm is
written in C-code, and is applied in real-time via d-space control desk on a windows
based PC, (this interface is shown in fig. F.2). 3-phase supply is fed into the CW-IPM
machine through a power analyser, (which measures the input quantities to the motor, as
well as the power factor). The CW-IPM machine is loaded by the Kollmorgen machine,
connected to a load bank. At low speeds, the Kollmorgen machine is connected to 4:1
gearbox to achieve higher loading torque. Output shaft torque from the CW-IPM
machine is measured by the 10Nm HBM torque transducer.
Fig. F.2 d-space control desk, real-time graphical user interface
Appendix F
264
F.2 Control Algorithm
The key portion of the control algorithm, written C, and ran with d-space control desk is
shown below:
/ ****************************************************************************************************
* Rot or Fi el d Or i ent ed Cont r ol of t he CW- I PM ( DS1104) *
****************************************************************************************************/
#i ncl ude <c: \ dSPACE\ DS1104\ RTLi b\ br t env. h>
#i ncl ude <ct r l . h> / * gener al header f or mot or cont r ol */
#i ncl ude <vl i mt f un. c>
#i ncl ude <svmodc. c>
/ * var i abl es f or execut i on t i me measur ement */
i nt i ndex_i nc;
i nt cnt ;
i nt i ndex_i nc;
Fl oat 64 exec_t i me;
Fl oat 64 i q_r ef 1;
Fl oat 64 i d_r ef 1;
Fl oat 64 v_oc;
Fl oat 64 Van;
Fl oat 64 I aExt ;
Fl oat 64 scal ei d=1;
Fl oat 64 speed_w_base=53. 0;
Fl oat 64 v_om=Vam+I am*Rs;
Fl oat 64 speed_w_cor ner =69. 0/ Pp/ 0. 168;
Fl oat 64 i d_man=0;
/ * adj ust val ues f or t i mer 0 */
Fl oat 64 t i mer 0_per i od = T_S;
/ * var i abl es f or communi cat i on wi t h Sl ave DSP */
I nt 16 t ask_i d = 0; / * communi cat i on channel */
I nt 16 i ndex = - 1; / * sl ave DSP command i ndex */
/ * par amet er s f or PWM i ni t i al i zat i on */
Fl oat 64 pwm_per i od = T_S; / * PWM per i od */
Fl oat 64 deadband = 2e- 6; / * deadband per i od */
UI nt 16 sync_mode = SLVDSP1104_PWM3_SYNC_CENT;
/ * i nt er r upt ser vi ce r out i ne t i mer 0 */
voi d i sr _t i mer 0( voi d)
{
/ * over r un check, enabl e i nt er r upt s gl obal l y */
ds1104_begi n_i sr _t i mer 0( ) ; / * st ar t i nt er r upt ser vi ce r out i ne t i mer */
RTLI B_TI C_START( ) ; / * st ar t t i me measur ement */
host _ser vi ce( 1, 0) ; / * Dat a Acqui si t i on ser vi ce */
/ ************************Rot or Posi t i on and Speed Acqui si t i on***************************/
psn_i ni t = psn_i ni t 1*PI / 180. 0;
angl e_l i mi t ( psn_i ni t ) ;
/ *r ead wi t h hi ghest r esol ut i on, 1/ 4 l i ne */
i nc_k = ds1104_i nc_count er _r ead( 1) ;
eps_m1 = eps_m1+( doubl e) ( i nc_k - i nc_k1) *PI 2/ 4. 0/ 5000. 0;
eps_m=eps_m1 + psn_i ni t ;
i ndex_i nc = ds1104_i nc_i ndex_r ead( 1, DS1104_I NC_I DXMODE_ON) ;
angl e_l i mi t ( eps_m) ;
/ *cal cul at e el ect r i cal angl e bet ween t he r ot or and st at or */
/ / Dan' s code t o l i mi t qngl e t o no mor e t han PI
eps_r s = Pp * eps_m;
i f ( eps_r s>6*PI ) eps_r s=eps_r s- 6*PI ;
el se i f ( eps_r s<=- 6*PI ) eps_r s=eps_r s+6*PI ;
el se i f ( eps_r s>5*PI ) eps_r s=eps_r s- 6*PI ;
el se i f ( eps_r s<=- 5*PI ) eps_r s=eps_r s+6*PI ;
el se i f ( eps_r s>4*PI ) eps_r s=eps_r s- 4*PI ;
el se i f ( eps_r s<=- 4*PI ) eps_r s=eps_r s+4*PI ;
el se i f ( eps_r s>3*PI ) eps_r s=eps_r s- 4*PI ;
el se i f ( eps_r s<=- 3*PI ) eps_r s=eps_r s+4*PI ;
el se i f ( eps_r s>2*PI ) eps_r s=eps_r s- 2*PI ;
el se i f ( eps_r s<=- 2*PI ) eps_r s=eps_r s+2*PI ;
el se i f ( eps_r s>PI ) eps_r s=eps_r s- 2*PI ;
el se i f ( eps_r s<=- PI ) eps_r s=eps_r s+2*PI ;
Appendix F
265
eps_r s_k = eps_m;
i f ( ( eps_r s_k<eps_r s_k1- 5000*PI / 30*T_S) | | ( eps_r s_k>eps_r s_k1+5000*PI / 30*T_S) ) {}
el se {ch_eps_r s = eps_r s_k - eps_r s_k1; }
eps_r s_k1 = eps_r s_k;
speed_w = ch_eps_r s / T_S;
f ol ( speed_w, speed_w_f ol , f ol _speed) ;
speed_spd = speed_w * 60. 0 / ( PI 2) ;
f ol ( speed_spd, speed_spd_f ol , f ol _spd) ;
i nc_k1 = i nc_k;
/ **************************DC Bus Vol t age and Cur r ent Acqui si t i on***********************/
/ *V_DC measur ement */
ds1104_adc_mux( 1) ;
ds1104_adc_del ayed_st ar t ( DS1104_ADC1) ;
V_DC = SCALE_VOLTAGE * ds1104_adc_r ead_ch( 1) ;
/ * get phase cur r ent s f r omADCs */
ds1104_adc_st ar t ( DS1104_ADC2 | DS1104_ADC3 | DS1104_ADC4 | DS1104_ADC5 ) ;
i _abc. a = SCALE_CURRENT * ds1104_adc_r ead_ch( 7) ;
i _abc. c = SCALE_CURRENT * ds1104_adc_r ead_ch( 6) ;
i _abc. b = - i _abc. a - i _abc. c;
Van= ds1104_adc_r ead_ch( 8) ;
I aExt = ds1104_adc_r ead_ch( 5) ;
i a = i _abc. a;
i b = i _abc. b;
i c = i _abc. c;
/ * conver t phase cur r ent s t o st at or cur r ent vect or */
f c3_2_spec( i _abc, i _ab) ;
/ *dq- axes cur r ent */
si ncos( eps_r s, t r i g_f ct ) ;
vr _neg( i _ab, t r i g_f ct , i _dq) ;
i d=i _dq. d;
i q=i _dq. q;
i d=i _ab. al pha*cos( eps_r s) +i _ab. bet a*si n( eps_r s) ;
i q=- i _ab. al pha*si n( eps_r s) +i _ab. bet a*cos( eps_r s) ;
/ ******************************PI Speed Cont r ol l er ******************************/
count ++;
i f ( count >= 5)
{
count = 0;
u_er r or _spd = speed_r ef - speed_w_f ol ;
u_er r or _p = Kp_spd*u_er r or _spd;
u_er r or _i = Ki _spd*u_er r or _spd+K_ant i _spd*( i q_r ef - u_out _spd) +u_er r or _i ;
i f ( u_er r or _i >= i q_max) { u_er r or _i = i q_max; }
i f ( u_er r or _i <=- i q_max) { u_er r or _i =- i q_max; }
u_out _spd=u_er r or _p+u_er r or _i ;
i q_r ef =u_out _spd;
i q_max=I am;
i f ( i q_r ef >= i q_max) { i q_r ef = i q_max; }
i f ( i q_r ef <=- i q_max) { i q_r ef =- i q_max; }
}
/ ********************************************************************************************/
/ ************************** Det er mi nat i on of Ref er ence and Max i d ***************************/
/ ********************************************************************************************/
i d_max = - 4;
i d_r ef =i d_man;
i f ( i d_r ef <= i d_max) {i d_r ef = i d_max; }
/ ********************************************************************************************/
Appendix F
266
/ ******************************PI Cur r ent Cont r ol l er s******************************/
/ / i d cont r ol l er
i d_er r or =i d_r ef - i d;
i d_er r or _p = Kp_d*i d_er r or ;
i d_er r or _i = Ki _d*i d_er r or +K_ant i _d*( vd_r ef - vd_out ) +i d_er r or _i ;
i f ( i d_er r or _i >= 0. 866*V_DC) { i d_er r or _i = 0. 866*V_DC; }
i f ( i d_er r or _i <=- 0. 866*V_DC) { i d_er r or _i =- 0. 866*V_DC; }
vd_out =i d_er r or _p+i d_er r or _i - ( 0. 0*Pp*speed_w_f ol *Lq*i q) ; / / Decoupl i ng t er ms i n br acket s
vd_r ef =vd_out ;
i f ( vd_r ef >= 0. 866*V_DC) { vd_r ef = 0. 866*V_DC; }
i f ( vd_r ef <=- 0. 866*V_DC) { vd_r ef =- 0. 866*V_DC; }
/ / i q cont r ol l er
i q_er r or =i q_r ef - i q;
i q_er r or _p = Kp_q*i q_er r or ;
i q_er r or _i = Ki _q*i q_er r or +K_ant i _q*( vq_r ef - vq_out ) +i q_er r or _i ;
i f ( i q_er r or _i >= 0. 866*V_DC) { i q_er r or _i = 0. 866*V_DC; }
i f ( i q_er r or _i <=- 0. 866*V_DC) { i q_er r or _i =- 0. 866*V_DC; }
vq_out =i q_er r or _p+i q_er r or _i +( 0. 0*Pp*speed_w_f ol *( FLUXM+Ld*i d) ) ; / / Decoupl i ng t er ms i n br acket s
vq_r ef =vq_out ;
i f ( vq_r ef >= 0. 866*V_DC) { vq_r ef = 0. 866*V_DC; }
i f ( vq_r ef <=- 0. 866*V_DC) { vq_r ef =- 0. 866*V_DC; }
/ *************************Decoupl i ng & Par k' s Tr ansf or mat i on***************************/
/ / Decoupl i ng
vd=vd_r ef ;
vq=vq_r ef ;
Ual pha=vd*cos( eps_r s) - vq*si n( eps_r s) ;
Ubet a =vq*cos( eps_r s) +vd*si n( eps_r s) ;
/ **************************************SVM*****************************************/
/ *vol t age l i mi t at i on*/
sv_l i m( &Uab, Ual pha, Ubet a, V_DC) ;
Ua=Uab. al pha;
Ub=Uab. bet a;
/ *space vect or modul at i on*/
sv_mod( &dut yCycl e, &sv_pwm, Uab. al pha, Uab. bet a, V_DC) ;
t 1=pwm_per i od*sv_pwm. t 1;
t 2=pwm_per i od*sv_pwm. t 2;
/ ************************************************************************************/
exec_t i me = RTLI B_TI C_READ( ) ;
/ * over r un check end, di sabl e i nt er r upt s gl obal l y */
ds1104_end_i sr _t i mer 0( ) ;
}
/ * i nt er r upt ser vi ce r out i ne f or PWM sync i nt er r upt */
voi d PWM_sync_i nt er r upt ( voi d)
{
/ * wr i t e PWM3 dut y cycl e t o sl ave DSP and t est f or er r or */
er r or = ds1104_sl ave_dsp_pwm3sv_dut y_wr i t e( t ask_i d, i ndex,
sv_pwm. sect or , t 1, t 2) ;
i f ( er r or ! = DSCOMDEF_NO_ERROR)
{
wr i t e_PWM3sv_er r or =1;
}
}
mai n( )
{
i ni t ( ) ;
/ * i nput si gnal f or channel 1 vi a TTL */
ds1104_i nc_i ni t ( 1, DS1104_I NC_MODE_RS422) ;
ds1104_i nc_set _i dxmode( 1, DS1104_I NC_I DXMODE_ON) ;
/ * i ni t i ncr ement al encoder channel 1 */
i ndex_i nc = 0;
whi l e ( i ndex_i nc == 0)
{
Appendix F
267
i ndex_i nc = ds1104_i nc_i ndex_r ead( 1, DS1104_I NC_I DXMODE_ON) ;
}
ds1104_i nc_count er _cl ear ( 1) ;
/ * i ni t i al i zat i on of sl ave DSP communi cat i on */
ds1104_sl ave_dsp_communi cat i on_i ni t ( ) ;
/ * i ni t and st ar t of 3- phase PWMSV gener at i on on sl ave DSP */
ds1104_sl ave_dsp_pwm3sv_i ni t ( t ask_i d, pwm_per i od, sv_pwm. sect or , t 1, t 2, deadband, sync_mode) ;
ds1104_sl ave_dsp_pwm3_st ar t ( t ask_i d) ;
/ * r egi st r at i on of PWM dut y cycl e updat e command */
ds1104_sl ave_dsp_pwm3sv_dut y_wr i t e_r egi st er ( t ask_i d, &i ndex) ;
/ * i ni t i al i zat i on of PWM sync i nt er r upt */
ds1104_set _i nt er r upt _vect or ( DS1104_I NT_SLAVE_DSP_PWM, ( DS1104_I nt _Handl er _Type)
&PWM_sync_i nt er r upt , SAVE_REGS_ON) ;
ds1104_enabl e_har dwar e_i nt ( DS1104_I NT_SLAVE_DSP_PWM) ;
RTLI B_I NT_ENABLE( ) ;
/ * i ni t i al i ze cont r ol par amet er s */
var i ni t ( ) ;
/ * per i odi c event wi t h t i mer 0 */
ds1104_st ar t _i sr _t i mer 0( t i mer 0_per i od, i sr _t i mer 0) ;
/ * Backgr ound t asks */
whi l e( 1)
{
RTLI B_BACKGROUND_SERVI CE( ) ; / * Cont r ol Desk ser vi ce */
}
}
/*********************************************** End ********************************************************/
F.3 3-phase IGBT Inverter
The 3-phse IGBT inverter which supplies power to the CW-IPM machine is shown fig.
F.3. Details of its circuit diagrams are shown in fig. F.4 to fig. F.5.
Fig. F.3 3-phse IGBT inverter (casing off) [174]
Appendix F
268
Fig. F.4 IGBT inverter schematic
Appendix F
269
Fig. F.5 Connections between the IGBT inverter and control boards
Appendix F
270
F.4 Kollmorgen PM Machine Specifications
Specifications of the Kollmorgen machine, (used as the loading machine), are as listed
in table F.1.
Table F.1
Kollmorgen PM Machine specifications
Model AKM33H
Rated Current 5.63Arms
Rated Torque 2.88Nm
Rated Voltage 320V
DC
Rated Speed 5500RPM
Rated Power 1.31kW
Stator Resistance 1.96O
(l-l)
Appendix G
271
APPENDIX G
PUBLICATION LIST
Patents:
[1] M. F. Rahman, Rukmi Dutta, Lester Chong Interior permanent magnet machine,
Provisional patent no. : 2011903320
Journal Publications:
[2] Lester Chong, Rukmi Dutta, M. F. Rahman, Application of Concentrated Windings in
Interior Permanent Magnet Machine, Australian J ournal of Electrical & Electronics
Engineering (AJ EEE) 2008. ISSN: 1448837X
[3] Lester Chong, M. F. Rahman, Saliency Ratio Derivation and Optimization for an IPM
Machine with Concentrated Windings Using Finite Element Analysis, Institution of
Engineering and Technology (IET) 2009. ISSN: 1751-8660
[4] Lester Chong, Rukmi Dutta, M. F. Rahman, Electromagnetic Losses in a 1kW
Concentric Wound IPM Machine for Field Weakening Applications, J ournal of
Applied Superconductivity and Electromagnetics (J ASEM), 2010. ISSN 1836-7151
Conference Publications:
[5] Lester Chong, Rukmi Dutta, M. F. Rahman, Application of Concentrated Windings in
Interior Permanent Magnet Machine, Australasian Universities Power Engineering
Conference (AUPEC) 2007, Australia. ISBN: 978-0-646-49488-3
[6] Lester Chong, Rukmi Dutta, M. F. Rahman, Open Circuit Analysis of Concentrated
Winding in Interior Permanent Magnet Machines with Fractional Slot Distribution, 4th
IET International Conference on Power Electronics, Machines and Drives (PEMD)
2008, UK. ISBN: 978-0-86341-900-3
[7] Lester Chong, M. F. Rahman, Comparison of d- and q-axis Inductances in an IPM
machine with Integral-slot Distributed and Fractional-slot Concentrated Windings, 18th
International Conference on Electrical Machines (ICEM) 2008, Portugal. ISBN: 978-1-
4244-1735-3
[8] Lester Chong, M. F. Rahman, Saliency Ratio Optimization in an IPM Machine with
Fractional-slot Concentrated Windings 11th International Conference on Electrical
Machines and Systems (ICEMS) 2008, China. ISBN: 978-1-4244-3826-6
[9] Lester Chong, Rukmi Dutta, M. F. Rahman, Parameter Analysis of an IPM Machine
with Fractional-slot Concentrated Windings, Part I: Open-circuit Analysis, Australasian
Universities Power Engineering Conference (AUPEC) 2008, Australia. ISBN: 978-0-
7334-2715-2
[10] Lester Chong, Rukmi Dutta, M. F. Rahman, Parameter Analysis of an IPM Machine
with Fractional-slot Concentrated Windings, Part II: Including Armature-reaction,
Australasian Universities Power Engineering Conference (AUPEC) 2008, Australia.
ISBN: 978-0-7334-2715-2
[11] Lester Chong, Rukmi Dutta, M. F. Rahman, Design of IPM machine with Concentrated
Windings for Vehicular Applications, European Power Engineering Conference, 2009,
Spain. ISBN: 978-1-4244-4432-8
Appendix G
272
[12] Lester Chong, Rukmi Dutta, M. F. Rahman, Design and Mechanical Consideration of
an IPM Machine with Concentrated Windings, Australasian Universities Power
Engineering Conference (AUPEC) 2009, Australia. ISBN: 978-1-4244-5153-1
[13] Lester Chong, Rukmi Dutta and M. F. Rahman, "Design of an interior permanent
magnet machine with concentrated winding for field weakening applications," in Proc.
of IEEE Int. Electric Machines & Drives Conf. (IEMDC), 2009, pp. 1985-1992.
[14] Lester Chong, Rukmi Dutta, M. F. Rahman, Design and Thermal Considerations of an
Interior Permanent Magnet Machine with Concentrated Windings, International
Conference on Electrical Machines and Systems (ICEMS) 2009, J apan. ISBN: 978-1-
4244-5177-7
[15] Lester Chong, Rukmi Dutta, M. F. Rahman, Field Weakening Performance of a
Concentrated Wound PM Machine with Rotor and Magnet Geometry Variation, Power
and Engineering Society General meeting (PES) 2010, USA. ISBN: 978-1-4244-6549-1
[16] Lester Chong, Rukmi Dutta, M. F. Rahman, Design of a Highly Efficient 1kW IPM
Machine with a Very Wide Constant Power Speed Range, International Power
Electronics Conference -ECCE-Asia (IPEC) 2010, J apan. ISBN: 978-1-4244-5394-8
[17] Lester Chong, Rukmi Dutta, M. F. Rahman, A Comparative Study of Rotor Losses in
an IPM with Single and Double Layer Concentrated Windings, International
Conference on Electrical Machines and Systems (ICEMS) 2010, Korea. ISBN: 978-1-
4244-7720-3
[18] Lester Chong, Rukmi Dutta, Howard Lovatt, Nguyen Quang Dai, M. F. Rahman,
Comparison of Concentrated and Distributed Windings in an IPM Machine for Field
Weakening Applications, Australasian Universities Power Engineering Conference
(AUPEC) 2010, Australia. ISBN: 978-1-4244-8379-2
[19] Lester Chong, Rukmi Dutta, M. F. Rahman, Howard Lovatt Experimental verification
of Rotor Losses in a Concentrated Wound IPM Machine with V-Shaped magnets,
International Electrical Machines and Drives Conference (IEMDC) 2011, Canada
[20] Rukmi Dutta, Lester Chong, M. F. Rahman, Analysis of CPSR in Motoring and
Generating Modes of an IPM Motor, International Electrical Machines and Drives
Conference (IEMDC) 2011, Canada
[21] Lester Chong, Rukmi Dutta, M. F. Rahman, Howard Lovatt Open Circuit Analysis of
an IPM Machine with Concentrated Windings Including Experimental Verification,
International Conference on Electrical Machines and Systems (ICEMS) 2011, China
[22] Lester Chong, Rukmi Dutta, D. Xiao, M. F. Rahman Performance Comparison between
Concentrated and Distributed Wound IPM Machines used for Field Weakening
Applications, Aegean Conference on Electric Machines and Power Electronics
(ACEMP) 2011, Turkey

Potrebbero piacerti anche