Design Optimization of Permanent Magnet Machines Over A Target Op PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 296

Marquette University

e-Publications@Marquette

Dissertations, Theses, and Professional


Dissertations (1934 -) Projects

Design Optimization of Permanent Magnet Machines Over a


Target Operating Cycle Using Computationally Efficient
Techniques
Alireza Fatemi
Marquette University

Follow this and additional works at: https://epublications.marquette.edu/dissertations_mu

Part of the Electrical and Computer Engineering Commons

Recommended Citation
Fatemi, Alireza, "Design Optimization of Permanent Magnet Machines Over a Target Operating Cycle
Using Computationally Efficient Techniques" (2016). Dissertations (1934 -). 662.
https://epublications.marquette.edu/dissertations_mu/662
DESIGN OPTIMIZATION OF PERMANENT MAGNET
MACHINES OVER A TARGET OPERATING
CYCLE USING COMPUTATIONALLY
EFFICIENT TECHNIQUES

By
Alireza Fatemi, A.S., B.S., M.S.

A Dissertation Submitted to the Faculty of the


Graduate School, Marquette University,
in Partial Fulfillment of the Requirements for the
Degree of Doctor of Philosophy

Milwaukee, Wisconsin
August 2016
ABSTRACT
DESIGN OPTIMIZATION OF PERMANENT MAGNET
MACHINES OVER A TARGET OPERATING
CYCLE USING COMPUTATIONALLY
EFFICIENT TECHNIQUES

Alireza Fatemi, A.S., B.S., M.S.


Marquette University, 2016

The common practices of large-scale finite element (FE) model-based design


optimization of permanent magnet synchronous machines (PMSMs) oftentimes
aim at improving the machine performance at the rated operating conditions,
thus overlooking the performance treatment over the entire range of operation
in the constant torque and extended speed regions. This is mainly due to the
computational complexities associated with several aspects of such large-scale design
optimization problems, including the FE-based modeling techniques, large number
of load operating points for load-cycle evaluation of the design candidates, and large
number of function evaluations required for identification of the globally optimal
design solutions.
In this dissertation, the necessity of accommodating the entire range of operation
in the design optimization of PMSMs is demonstrated through joint application of
numerical techniques and mathematical or statistical analyses. For this purpose,
concepts such as FE analysis (FEA), design of experiments (DOE), sensitivity
analysis, response surface methodology (RSM), and regression analysis are extensively
used throughout this work to unscramble the correlations between various factors
influencing the design of PMSMs. Also in this dissertation, computationally efficient
methodologies are developed and employed to render unprohibitive the problems
associated with large-scale design optimization of PMSMs over the entire range of
operation of such machines. These include upgrading an existing computationally
efficient FEA to solve the electromagnetic field problem at any load operating point
residing anywhere in the torque-speed plane, developing a new stochastic search
algorithm for effectively handling the constrained optimization problem (COP) of
design of electric machines so as to reduce the number of function evaluations required
for identifying the global optimum, implementing a k-means clustering algorithm for
efficient modeling of the motor load profile, and devising alternative computationally
efficient techniques for calculation of strand eddy current losses or characterization of
the mechanical stress due to the centrifugal forces on the rotor bridges.
The developed methodologies in this dissertation are applicable to the wide class
of sine-wave driven PM and synchronous reluctance machines. Here, they were
successfully utilized for optimization of two existing propulsion traction motors over
predefined operating cycles. Particularly, the well-established benchmark design
provided by the Toyota Prius Gen. 2 V-type interior PM (IPM) motor, and a
challenging high power density spoke-type IPM for a formula E racing car are treated.
i

ACKNOWLEDGEMENTS

Alireza Fatemi, A.S., B.S., M.S.

For any accomplishment, including my doctorate degree, I am grateful beyond


words to my parents, Mr. Mohammad Fatemi and Mrs. Meymanat Shadkam,
who gave me the gift of life, devoted their own lives to others’ well-being, trusted
in my abilities, encouraged me to have strong work ethics, and set me up for this
journey. Yet, this dissertation would have not been possible without the guidance,
encouragement, and support of a great number of other individuals whom I had the
privilege to work with and learn from during the past four years, and to whom I will
be indebted for the years to come.
First and foremost, I am thankful to Prof. Nabeel Aly Omar Demerdash, who, as
an unparalleled academic advisor and pedagogue, supported me above and beyond
that of any professor, throughout my Ph.D. studies, and mentored my scholarly, and
intellectual development in numerous ways. The countless hours that he spent with
me in person to teach me computational electromagnetics, discuss my research, read
and evaluate my papers and dissertation word for word, and tell invaluable life stories
and history lessons will for sure remain the most memorable moments of these years.
For all he has done for me, Prof. Demerdash has expected nothing from me but to
be as caring and supportive to my own students, should I choose an academic career
path in the future, to whom I pledge my personal commitment.
I am very grateful to Prof. Dan M. Ionel, of the Department of Electrical and
Computer Engineering of the University of Kentucky, for his instrumental role in
every aspect of the research presented in this dissertation. In addition to his personal
and close involvement, he established invaluable collaborations with renowned experts
in this industry to pursue pragmatic research and to bridge academic scholarship and
industrial practice. The depth and breadth of this work owes itself to Prof. Ionel’s
never-ending zest for research and its dissemination.
I am gratefully indebted to Dr. Thomas W. Nehl, of General Motors (GM) Global
Research and Development, for all he did during these years from initiating the drive-
cycle optimization project which lent itself to becoming the centerpiece of my Ph.D.
research and dissertation, to providing the internship opportunity in his reputable
team which led to starting my career at GM. His support has been essential during
this time, and so have been his many contributions to the technical aspect of this
work. I would have not embarked on this project, had it not been for Dr. Nehl’s
entrepreneurial initiatives.
I would like to also acknowledge the generous participation of Prof. Thomas M.
Jahns, of the Department of Electrical and Computer Engineering of the University
of Wisconsin-Madison, and Prof. Edwin E. Yaz and Prof. James E. Richie, of the
Department of Electrical and Computer Engineering of Marquette University, in my
Ph.D. Dissertation Committee. Particularly, Prof. Jahn’s insightful commentary
prompted mind-provoking discussions and substantially enriched the content of this
dissertation, for which I am very much thankful.
ii

I would like to express my special thanks to my dear friends and senior members of
the Electric Machines and Drives Lab (EMDL) at Marquette University, Dr. Gennadi
Y. Sizov, of Rockwell Automation, and Dr. Peng Zhang, of General Motors, for
laying a solid foundation for the automated design optimization of permanent magnet
machines, which served as the start point of my dissertation research. Their support
in the initial stages of this project, and their encouragement in its continuation was
exceptional.
Throughout my Ph.D. studies, I had the opportunity to collaborate with and
learn from many experts in the field. During the early stages of this project, Mr.
James R. Hendershot’s instructions in recent advances in the design of brushless PM
machines boosted my confidence to delve into this topic. Likewise, the instructive role
of Dr. Dave A. Staton and Dr. Mircea Popescu, of Motor Design Ltd., in multiphysics
analysis of PM machines was of great help and should be acknowledged. I also learned
from, and enjoyed very much the collaborations with Mr. Steven J. Stretz, of Regal
Beloit Company, and Dr. Rafal Wrobel and Dr. Yew C. Chong, of Motor Design Ltd.
I am also thankful to Mr. Mark G. Solveson, of Ansys Inc., for his technical support
for advanced use of Ansys software packages.
I would like to also thank the members of the Electric Drives and Power Electronics
Systems Lab at GM R&D including Dr. Lei Hao, Dr. Chandra S. Namuduri, Dr.
Rashmi Prasad, Dr. Suresh Gopalakrishnan, and Dr. Avoki Omekanda, for their
mentorship and collegiality during summer 2015.
I would like to acknowledge the financial support of Marquette University’s EECE
Department, GM, M-WERC Consortium, and SEPMEED Consortium, which allowed
me to work in this exciting and challenging field. The software support of Ansys Inc.,
and Motor Design Ltd. is also gratefully acknowledged.
I am also thankful to my present and former colleagues and friends at Marquette
University’s Electric Machines and Drives Lab, Dr. Gennadi Sizov, Dr. Peng Zhang,
Dr. Jiangbiao He, Mr. Chad Somogyi, Mr. Andrew Strandt, Ms. Alia Strandt, and
Mr. Muyang Li for creating a genuinely collegial learning and work environment.
Also many thanks to my friends, Mr. Ahmadreza Baghaie, Dr. Benyamin Davaji,
Dr. Marek Trawicki, Dr. Ahmed Sayed Ahmed, and Mr. Ramin Katebi for their
encouragement and support.
Last but surely not least, I am eternally grateful to my most faithful companion,
Nancy, whose love has inspired my life, and who is my home away from home.

Alireza Fatemi
August, 2016
iii

TABLE OF CONTENTS

ACKNOWLEDGEMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . i

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x

1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1 Significance of the Research Topic . . . . . . . . . . . . . . . . . . . . 1

1.2 Design Optimization of PM Machines Based on Computational


Electromagnetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.2.1 Early Developments . . . . . . . . . . . . . . . . . . . . . . . 5

1.2.2 Design Optimization for Rated Operation . . . . . . . . . . . 10

1.2.3 Design Optimization for Entire Operating Range . . . . . . . 14

1.3 Statement of the Problem . . . . . . . . . . . . . . . . . . . . . . . . 19

1.4 Dissertation Organization . . . . . . . . . . . . . . . . . . . . . . . . 21

1.5 Related Publications . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2 DESIGN OPTIMIZATION FOR RATED OPERATING CONDI-


TIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.1 Design Synthesis Using CE-FEA and DE . . . . . . . . . . . . . . . . 26

2.1.1 Overview of Computationally Efficient-Finite Element Analysis 26

2.1.2 Overview of Differential Evolution . . . . . . . . . . . . . . . . 35


iv

2.2 Case Study Design Optimization of Toyota Prius Gen 2 Motor at Its
Nominal Operating Point . . . . . . . . . . . . . . . . . . . . . . . . . 39

2.2.1 Parametric CE-FEA Model . . . . . . . . . . . . . . . . . . . 40

2.2.2 Optimization Procedure and Results . . . . . . . . . . . . . . 45

2.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

3 EFFECTS OF AMPERE LOADING LEVEL ON OPTIMAL


DESIGN OF IPM MOTORS . . . . . . . . . . . . . . . . . . . . . . . 61

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

3.2 Benchmark Designs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

3.2.1 Parametrized Cross Sections . . . . . . . . . . . . . . . . . . . 65

3.2.2 Time-Stepping FE Models . . . . . . . . . . . . . . . . . . . . 67

3.3 Parallel Sensitivity Analysis at Different Ampere Loading Levels . . . 70

3.3.1 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

3.3.2 Discussion of the Results of the Sensitivity Analysis . . . . . . 71

3.4 Parallel Statistical Analysis of the Optimized Designs at Different


Ampere Loading Levels . . . . . . . . . . . . . . . . . . . . . . . . . . 76

3.4.1 Large-Scale Design Optimization . . . . . . . . . . . . . . . . 77

3.4.2 Scaling Rules of the Optimum Candidate Designs . . . . . . . 79

3.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

4 MULTIOBJECTIVE CMODE-TYPE OPTIMIZATION OF ELEC-


TRIC MACHINES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

4.2 CMODE Versus DE in the Design Optimization of Electric Machines 89

4.3 Benchmark Study- Application to IPM Motors . . . . . . . . . . . . . 94


v

4.3.1 Parametrized FE Model . . . . . . . . . . . . . . . . . . . . . 94

4.3.2 Optimization Fitness Functions . . . . . . . . . . . . . . . . . 98

4.4 Comparative Study of the Results . . . . . . . . . . . . . . . . . . . . 100

4.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

5 DRIVE-CYCLE PERFORMANCE OPTIMIZATION . . . . . . . 115

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

5.2 Efficient Modeling of the Motor Load Profile . . . . . . . . . . . . . . 119

5.2.1 Energy Distribution Over the Torque-Speed Plane . . . . . . . 119

5.2.2 Cyclic Representative Points . . . . . . . . . . . . . . . . . . . 124

5.3 Performance Evaluation at Representative Points . . . . . . . . . . . 128

5.4 Extended Speed Operation of PM Machines . . . . . . . . . . . . . . 136

5.5 Drive-Cycle Optimization . . . . . . . . . . . . . . . . . . . . . . . . 145

5.5.1 Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145

5.5.2 Computational Complexity of the Optimization Algorithm . . 149

5.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

6 CASE STUDY DRIVE-CYCLE OPTIMIZATION OF TRACTION


MOTORS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

6.2 Optimization of the Prius IPM Motor Over a Compound Operating


Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

6.2.1 Initialization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

6.2.2 Optimization Results . . . . . . . . . . . . . . . . . . . . . . . 157


vi

6.3 Optimization of a Formula E Racing Car IPM Motor Over the Le Mans
Operating Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166

6.3.1 Initialization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168

6.3.2 Optimization Results . . . . . . . . . . . . . . . . . . . . . . . 176

6.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

7 ADDITIONAL IMPLICATIONS OF PM MACHINES’ DESIGN


OPTIMIZATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189

7.1 RSM-DE-ANN Sensitivity Analysis of Material Cost in PM Motors


with Distributed and Concentrated Windings . . . . . . . . . . . . . 191

7.1.1 FE-Based Machine Models . . . . . . . . . . . . . . . . . . . . 193

7.1.2 Effects of the Commodity Price on the Design Correlations . . 194

7.1.3 ANN-Based Design Optimization with Different Commodity


Price Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . 198

7.1.4 Sensitivity of Optimal Design Values to Commodity Price


Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202

7.2 A Computationally Efficient Method for Calculation of Strand Eddy


Current Losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206

7.2.1 Strand Eddy Current Loss Characterization . . . . . . . . . . 209

7.2.2 FE-Based Eddy Current Loss Estimation for Randomly Wound


Stator Windings . . . . . . . . . . . . . . . . . . . . . . . . . 211

7.2.3 Case-Study Analysis . . . . . . . . . . . . . . . . . . . . . . . 217

7.3 Estimation of Tangential Mechanical Stresses on the Rotor Bridges . 223

7.3.1 Adopted Analytical Stress Estimation Method . . . . . . . . . 225

7.3.2 Evaluation of the Accuracy of the Analytical Stress Estimation


Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227

7.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232


vii

8 CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235

8.1 Summary and Conclusion . . . . . . . . . . . . . . . . . . . . . . . . 235

8.2 Recommendation for Future Work . . . . . . . . . . . . . . . . . . . . 241

BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243

APPENDIX I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264

APPENDIX II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267

APPENDIX III . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272


viii

LIST OF TABLES

1.1 Estimated annual savings resulting from the utilization of high-


efficiency motor systems [1]. . . . . . . . . . . . . . . . . . . . . . . . 2

2.1 Boundaries of the design variables defined over the parameterized cross-
section of the Prius motor. . . . . . . . . . . . . . . . . . . . . . . . . 44

2.2 Geometric design parameters of the counterpart optimal designs


obtained from the Prius motor design optimization at nominal load
point. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

2.3 Comparison of nominal performance between counterpart designs


obtained from the Prius motor design optimization at nominal load
point. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

2.4 Distribution of power losses in the counterpart designs obtained from


the Prius motor design optimization at nominal load point. . . . . . . 51

2.5 Mass of active components in the counterpart designs obtained from


the Prius motor design optimization at nominal load point. . . . . . . 51

3.1 The independent design variables of the case study machines for
sensitivity analysis and optimization at different ampere loading levels. 67

3.2 Typical current density ranges found in electric machines with different
cooling systems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

3.3 Average design parameters of the optimized designs at different ampere


loading levels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

3.4 Mean of the ratio of copper losses to core losses in the Pareto-optimal
designs at different ampere loading levels. . . . . . . . . . . . . . . . . 82

3.5 Mean of the masses of the optimized designs normalized with respect
to the values obtained for NC class. . . . . . . . . . . . . . . . . . . . 83

4.1 Independent design variables of the parametric stator and rotor


structures shown in Figs. 4.3 and 4.4. . . . . . . . . . . . . . . . . . . 98
ix

4.2 Comparison of the number of the feasible Pareto optimal designs


between DE and CMODE. . . . . . . . . . . . . . . . . . . . . . . . . 112

4.3 Comparison of the number of total feasible design candidates between


DE and CMODE. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

6.1 Cyclic representative points for the combined US driving cycle shown
in Fig. 6.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

6.2 Cyclic representative points of Le Mans driving cycle shown in Fig. 6.13.171

6.3 Independent design variables and their upper and lower bounds of the
18-slot 16-pole spoke-type machine. . . . . . . . . . . . . . . . . . . . 172

6.4 The design characteristics of the counterpart spoke-type motors. . . . 180

6.5 The design characteristics of the optimized high power density “D3”
motor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
x

LIST OF FIGURES

1.1 U.S. electricity use by sector, (million kWh/year 2006) [2]. . . . . . . 1

1.2 Required improvements in current electric drive systems to meet EV


Everywhere requirements [3]. . . . . . . . . . . . . . . . . . . . . . . . 3

1.3 Design optimization based on GA. . . . . . . . . . . . . . . . . . . . . 11

1.4 High level design flow of an electric machine. . . . . . . . . . . . . . . 20

2.1 Design optimization of PM machines based on CE-FEA and DE. . . . 27

2.2 Mapping the field values between typical sister elements in the CE-
FEA method. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

2.3 Reconstruction of mid-tooth flux densities using the CE-FEA method. 29

2.4 Reconstruction of mid-yoke flux densities using the CE-FEA method. 30

2.5 Reconstruction of stator winding flux linkages using the CE-FEA method. 31

2.6 Reconstruction of the torque profile using the CE-FEA method. . . . 32

2.7 Evolution model of DE from one generation to another. Multiple


members of the current population can be processed simultaneously
to decrease the computation time. . . . . . . . . . . . . . . . . . . . . 38

2.8 Cross-section of the Toyota Prius Gen 2 IPM motor. . . . . . . . . . 40

2.9 Parametrized stator slot of the Prius motor. . . . . . . . . . . . . . . 42

2.10 Parametrized rotor layout of the Prius motor. . . . . . . . . . . . . . 42

2.11 The parameterized cross-section of the Prius motor comprising 10


independent design variables, see Table 2.1. . . . . . . . . . . . . . . 45

2.12 Range of variation of the design parameters defined over the Prius
motor topology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

2.13 Determination of current advanced angle for MTPA operation. . . . . 47


xi

2.14 Results of the optimization of the Prius motor at the nominal load
point expressed in terms of the designated objectives. . . . . . . . . . 49

2.15 Cross-sections and rated field plots of the three counterpart designs
obtained from the Prius motor design optimization at nominal load
point. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

2.16 Line to line back EMF of the counterpart designs obtained from the
Prius motor design optimization at nominal load point (a) G59M12,
(b) G20M62, and (c) Prius. . . . . . . . . . . . . . . . . . . . . . . . 54

2.17 Harmonic content of the line to line terminal voltage of the counterpart
designs obtained from the Prius motor design optimization at nominal
load point. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

2.18 Torque ripple of the counterpart designs obtained from the Prius motor
design optimization at nominal load point. . . . . . . . . . . . . . . . 56

2.19 Efficiency maps of the counterpart designs obtained from the Prius
motor design optimization at nominal load point. . . . . . . . . . . . 57

2.20 Load operating points of Toyota Prius propulsion PM motor obtained


from ADVISOR. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

2.21 Load operating points of Honda Insight propulsion PM motor obtained


from ADVISOR. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

3.1 The parametrized cross section of the case study machines for
sensitivity analysis and optimization at different ampere loading levels. 66

3.2 Normalized regression coefficients from the sensitivity study indicating


the effect of the design variables on, (a) active material cost, (b) copper
loss, (c) core loss, (d) minimum PM flux density, and (e) torque ripple
for the 48S8P machine. . . . . . . . . . . . . . . . . . . . . . . . . . . 72

3.3 Normalized regression coefficients from the sensitivity study indicating


the effect of the design variables on, (a) active material cost, (b) copper
loss, (c) core loss, (d) minimum PM flux density, and (e) torque ripple
for the 12S10P machine. . . . . . . . . . . . . . . . . . . . . . . . . . 73

3.4 Optimization results for the three case-study cooling systems. . . . . 78

3.5 The distribution of the design variables in the optimized designs for
the three case study cooling systems. . . . . . . . . . . . . . . . . . . 80
xii

3.6 Optimized cross sections derived based on the mean of the design
variables in the 500 Pareto-optimal designs listed in Table 3.3, for the
three current density levels. . . . . . . . . . . . . . . . . . . . . . . . 82

4.1 Flowchart of the steps of the DE optimization algorithm. . . . . . . . 90

4.2 Flowchart of the steps of the CMODE optimization algorithm. . . . 91

4.3 The parameterized stator structures used for constructing the example
IPM motors for comparison between DE and CMODE. . . . . . . . . 96

4.4 The parameterized rotor structures used for constructing the example
IPM motors for comparison between DE and CMODE. . . . . . . . . 97

4.5 The evolution of the optimization process using DE for the three case-
study motors under the first scenario. . . . . . . . . . . . . . . . . . 102

4.6 The evolution of the optimization process using CMODE for the three
case-study motors under the first scenario. . . . . . . . . . . . . . . . 103

4.7 The evolution of the optimization process using DE for the three case-
study motors under the second scenario. . . . . . . . . . . . . . . . . 104

4.8 The evolution of the optimization process using CMODE for the three
case-study motors under the second scenario. . . . . . . . . . . . . . 105

4.9 Feasible Pareto optimal designs of the two optimization algorithms for
scenario 1 of the fitness functions. . . . . . . . . . . . . . . . . . . . 106

4.10 Feasible Pareto optimal designs of the two optimization algorithms for
scenario 2 of the fitness functions. . . . . . . . . . . . . . . . . . . . 107

4.11 Convergence of the feasible design candidates for the three case-study
motors in terms of loss × AM C for scenario 1. . . . . . . . . . . . . 109

4.12 Convergence of the feasible design candidates for the three case-study
motors in terms of loss × ripple for scenario 2. . . . . . . . . . . . . 110

4.13 The normalized hypervolumes of the Pareto fronts generated by DE


and CMODE for the two scenarios of fitness functions. . . . . . . . . 111

4.14 Typical optimized cross-sections and the field plots of the studied
motors for the two scenarios of fitness functions. . . . . . . . . . . . . 113

5.1 The Toyota Prius Gen 2. motor load profiles. . . . . . . . . . . . . . 120


xiii

5.2 The Honda Insight Gen. 1 motor load profile. . . . . . . . . . . . . . 121

5.3 Toyota Prius Gen. 2 motor output energy over torque-speed plane. . 122

5.4 Honda Insight Gen. 1 motor output energy over torque-speed plane. . 123

5.5 Cyclic representative points of Toyota Prius Gen. 2 motor output


energy over torque-speed plane. . . . . . . . . . . . . . . . . . . . . . 126

5.6 Cyclic representative points of Honda Insight Gen. 1 motor output


energy over torque-speed plane. . . . . . . . . . . . . . . . . . . . . . 127

5.7 The best sum of the distances of the load points to their corresponding
cluster means versus the number of clusters. . . . . . . . . . . . . . . 129

5.8 Effects of saturation and cross-saturation in prediction of torque, and


induced voltages over the full range of excitation current. . . . . . . . 130

5.9 Developed method for derivation of the stator winding currents at every
load point for time-stepping magneto-static FEA. Optimal control is
ensured for constant torque and flux weakening operation. . . . . . . 133

5.10 Process of derivation of the excitation current for a typical load point. 134

5.11 PM demagnetization maps of a typical motor for PMs located with


respect to the rotor motion for motoring operation. . . . . . . . . . . 135

5.12 Sampled flux contours of the three example motors with equal rated
torque and rated current density. . . . . . . . . . . . . . . . . . . . . 138

5.13 Efficiency maps of the three example IPMs with infinite maximum
speeds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

5.14 Loss ratio maps of the three example IPMs with infinite maximum
speeds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

5.15 Current density maps of the three example IPMs with infinite
maximum speeds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

5.16 Q-axis current density maps of the three example IPMs with infinite
maximum speeds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

5.17 Negative d-axis current density maps of the three example IPMs with
infinite maximum speeds. . . . . . . . . . . . . . . . . . . . . . . . . . 144
xiv

5.18 Current density maps of the three example IPMs with infinite
maximum speeds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

5.19 The flowchart of steps of the overall optimization algorithm. . . . . . 147

6.1 Load profile of the Toyota Prius Gen. 2 IPM over a combined driving
cycle. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

6.2 Toyota Prius Gen. 2 motor output energy versus torque and speed for
the combined US driving cycles. . . . . . . . . . . . . . . . . . . . . . 154

6.3 Cyclic representative points with seven clusters for the combined US
driving cycle. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

6.4 Optimization results of Toyota Prius Gen .2 IPM motor over the
combined US driving cycles. . . . . . . . . . . . . . . . . . . . . . . . 158

6.5 Comparison of the selected designs obtained from drive-cycle


optimization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159

6.6 Efficiency maps of the selected designs obtained from drive-cycle


optimization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

6.7 Tested efficiency map of the Toyota Prius Gen. 2 IPM motor reported
by the research team at ORNL. . . . . . . . . . . . . . . . . . . . . . 161

6.8 Lumped thermal network model of the motor cooling system developed
in Motor-CAD. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162

6.9 The peak temperatures of the counterpart designs evaluated over US06
driving cycle. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

6.10 Von-Mises stress throughout the rotor structure of the counterpart


designs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

6.11 Formula E motor load profile for the Le Mans driving cycle. . . . . . 168

6.12 Formula E motor energy consumption in the torque-speed plane. . . . 169

6.13 Le Mans driving cycle representative points obtained from k-means


algorithm with seven clusters. . . . . . . . . . . . . . . . . . . . . . . 170

6.14 The parameterized FE model of the spoke-type PM motor, see Table


6.3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
xv

6.15 The results of optimization of the spoke-type IPM over 3,400 design
solutions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174

6.16 Correlation of performance metrics with mass and power losses in the
designs optimized for efficiency and high power density. . . . . . . . . 175

6.17 Distribution of the design parameters in the Pareto-optimal designs,


and their correlation with total mass and power losses. . . . . . . . . 177

6.18 Equivalent machine model used for analytical investigation of the


optimal design of spoke-type FSCWs. . . . . . . . . . . . . . . . . . . 178

6.19 Histogram of the distributions of (a) ksi , and (b) kwpm in the 100 Pareto-
optimal designs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179

6.20 Flux lines distributions of the optimized designs at 6 000 r/min under
rated load. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

6.21 Efficiency maps of the optimized spoke-type designs. . . . . . . . . . 183

6.22 Copper loss maps of the optimized spoke-type designs. . . . . . . . . 184

6.23 Core loss maps of the optimized spoke-type designs. . . . . . . . . . . 185

6.24 Identifying a design with higher power density and drive-cycle efficiency
than the original design. . . . . . . . . . . . . . . . . . . . . . . . . . 186

6.25 Flux lines distributions at 6000 r/min. . . . . . . . . . . . . . . . . . 186

6.26 Efficiency maps of the high power density spoke-type designs. . . . . 187

7.1 Variation of the price of constructive commodities of Nd-based PM


motors for three consecutive years leading the surge of the Nd price in
2011 [4, 5]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191

7.2 Influence of the design parameters on AMC in the two case-study motors.195

7.3 Change of the influences of the design parameters on AMC based on


the value of the commodity price coefficients. . . . . . . . . . . . . . . 196

7.4 Developed procedure for sensitivity analysis of the optimal design values.200

7.5 Estimation error of the ANN-based models when compared to FE-


based machine models. . . . . . . . . . . . . . . . . . . . . . . . . . . 201
xvi

7.6 Optimization results of the 48S8P motor configuration for a typical set
of commodity price coefficients. . . . . . . . . . . . . . . . . . . . . . 203

7.7 Optimization results of the 12S10P motor configuration for a typical


set of commodity price coefficients. . . . . . . . . . . . . . . . . . . . 204

7.8 The sensitivity of the optimal design parameters to variation of the


commodity price coefficients. . . . . . . . . . . . . . . . . . . . . . . . 206

7.9 Slot leakage and fringing flux in a typical open-slot FSCW PM machine.207

7.10 Alternative coil models for strand eddy current loss analysis. . . . . . 212

7.11 Radial and tangential components of the field sections in Fig. 7.10(c)
for a typical motor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214

7.12 Determination of sf,max and conductor positions by moving the slot


geometry over a grid of conductors. . . . . . . . . . . . . . . . . . . . 215

7.13 Slot fill factor and strand positions for example slot geometries as the
net slot area increases. . . . . . . . . . . . . . . . . . . . . . . . . . . 216

7.14 Reconstruction of the field harmonics from the sample points using
Delaunay triangulation method. . . . . . . . . . . . . . . . . . . . . . 217

7.15 Mapped flux on each individual strand. . . . . . . . . . . . . . . . . . 218

7.16 Distribution of strand eddy current losses under various loading levels. 219

7.17 Case-study investigation of strand eddy current losses. . . . . . . . . 220

7.18 Comparison of the accuracy of the loss calculation method over a wide
range of frequencies and loading conditions. . . . . . . . . . . . . . . 221

7.19 Estimation error of the computationally efficient method of calculation


of strand eddy current losses compared to the full-fledged time
harmonic analysis with detailed coil modeling. . . . . . . . . . . . . . 222

7.20 Variation of dc ohmic losses and strand eddy current losses with respect
to loading level. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223

7.21 Ratio of ac to dc losses over a wide range of loading conditions. . . . 224

7.22 Modeling of the original layout on the left with the equivalent ring on
the right. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
xvii

7.23 Preparing the rotor geometry for structural FE analysis. . . . . . . . 229

7.24 Comparison of the accuracy of the analytical stress estimation method


through a sensitivity analysis. . . . . . . . . . . . . . . . . . . . . . . 230

7.25 Change of kdpm ratio from lower to upper bound. . . . . . . . . . . . . 231

7.26 Change of kwq ratio from lower to upper bound. . . . . . . . . . . . . 231


1

CHAPTER 1
INTRODUCTION

1.1 Significance of the Research Topic

Electric motor driven systems are considered to be the largest consumers of electric

energy, often accounting for up to one-third or more of the total electricity sold in

industrialized countries [1]. In the U.S., electric motors account for 38% of the total

electrical energy consumption [2, 6], see Figure 1.1. The financial and environmental

incentives for energy savings are substantial, through improving the efficiency of the

electric motor driven systems in the U.S., see Table 1.1 [1].

In this respect, two fundamental areas of opportunity are:

• improving the motor efficiency at the component level, and

• improving the device efficiency at the system level.

Commercial,
498000, 13%

Other, Electric Motors,


2291000, 62% 1431000, 38%
Industrial, Residential,
632000, 17% 297000, 8%

Transport,
4000, 0%

Figure 1.1: U.S. electricity use by sector, (million kWh/year 2006) [2].
2

Table 1.1: Estimated annual savings resulting from the utilization of high-efficiency
motor systems [1].
Cost saved $3-5 billion
Energy saved 62-104 giga kWh
CO2 emission 15-26 mega tons

The latter measure, which will not be addressed in this dissertation, covers a

broad spectrum of strategies ranging from upgrading the system control technologies,

to matching the type and size of the motor with the load requirements, better

maintenance, etc.

Regarding the efficiency advancement of electric motors at the component level,

recent trends have focused on [7]:

• the use of new materials for laminations, conductors, and permanent magnets

(if applicable),

• the investigation of advanced concepts in design of the rotor, stator, and winding

configurations, and

• perfecting the existing technology by fully exploring the design space for optimal

sizing/shaping of the motor cross-section.

Owing to the Energy Policy and Conservation Act of 1992 (EPAct) and the

subsequent NEMA Premium motors standard, which has been made mandatory since

Dec. 2010, the penetration rate of high efficiency motors in the U.S. has been steadily

rising in recent years [2]. Lately, the EV Everywhere Grand Challenge [3], which

was initiated in March 2012 to render plug-in electric vehicles (PEVs) as affordable

as gasoline-powered vehicles, introduced a new category of stringent requirements on


3

4X Cost Reduction
35% Size Reduction
40% Weight Reduction
40% Loss Reduction

2022 Electric Drive System**


2012 Electric Drive System*
$8/kW, 1.4 kW/kg, 4 kW/L
$30/kW, 1.1 kW/kg, 2.6 kW/L
94% system efficiency
90% system efficiency
55kW SYSTEM cost of $440
55kW SYSTEM cost of $1650
** Future systems may meet these performance
* Today's electric drive systems use discrete
targets through advancements such as fully
components, silicon, semiconductors, and rare
integrating motors and electronics, wide bandgap
earth motor magnets.
semiconductors, and non-rare earth motor.

Figure 1.2: Required improvements in current electric drive systems to meet EV


Everywhere requirements [3].

electric drive systems in EV applications which should be realized by 2022, see Figure

1.2. Therefore, it is expected that performance enhancement of electric motors will

remain a continuous challenge in future years.

There are a number of factors by which high performance electric motors

are characterized and measured; amongst them are high efficiency, low torque

ripple, high torque density, high power factor, sinusoidal back-emf waveform with

low harmonic content, fast dynamics, mechanical robustness, low maintenance,

manufacturability, cost effectiveness, and sustainability. As found in research studies

by the American Council for an Energy-Efficient Economy (ACEEE) [8], the technical

difficulties of performance improvement in induction motors are prohibitive due


4

to their associated mature technologies, thus promoting further research in other

advanced commutatorless motors such as synchronous reluctance motors (SyRM),

permanent magnet synchronous motors (PMSM), flux switching motors (FSM),

switched reluctance motors (SRM), etc.

Due to the eliminated rotor copper losses in PMSMs, such machines can be

designed to maintain high efficiency at higher torque densities, and yet with a less

sophisticated cooling system [9]. With proper control, PMSMs can operate at high

power factors similar to their counterpart wound-field synchronous machines with

a separate controllable magnetic field source. However, as opposed to the wound-

field synchronous machines, PMSMs have lower maintenance, better rotor structural

integrity, and faster dynamics due to the absence of rotating field windings and

elimination of the brushes or other complex brushless excitation systems. These

characteristics have rendered PMSMs to represent a catagory of promising solutions

to meet the stringent requirements of high performance applications.

In this regard, this dissertation will focus on performance improvements in

sinewave drive brushless PMSM technology using computationally efficient and high

fidelity simulations as will be discussed in the following sections.


5

1.2 Design Optimization of PM Machines Based

on Computational Electromagnetics

1.2.1 Early Developments

Computational electromagnetics (CEM) began by applying the principles of highly

developed finite difference (FD) discipline of mathematics to solving two-dimensional

static field problems [10] in the works of Trutt [11], Erdelyi [12–15], and Demerdash

[16–18]. From the very beginning, computer aided design optimization of electric

machines based on CEM was pursued by many investigators. An example of

pioneering numerical design optimization of synchronous machines was reported

by Demerdash and Hamilton [16, 19], where finite difference models of two- and

four-pole turbo-generators were developed and employed for determining asymmetric

angular position of the rotor slots for reducing the eddy current losses in the stator

strands. The model was capable of taking the complex geometry and the saturation

phenomenon in to account.

As the finite element method (FEM) proved to be more effective, more accurate,

and easier to implement in electromagnetic modeling of electric machines [20, 21],

the tendency for application of FEM, as opposed to FD, in various types of electric

machines grew rapidly in early works of many investigators such as Chari and Silvester

[22–26], Brauer [27], Anderson [28], Demerdash and Nehl [21, 29–31], and others.

Some of the earliest FE-based design optimization efforts were conceived as

“inverse field problems”, as opposed to a “forward problem” in which the field


6

distribution and the magnetic vector potential is determined for given design

specifications including the geometry, excitation, boundaries, and material properties.

In an inverse problem, the flow of the design is reversed whereby a specific performance

objective such as flux distribution is pursued by treating the parameters previously

assumed to be known as unknowns [32, 33]. The first optimization attempts

formulated as inverse problems were achieved by handling the coordinates of the

boundary nodes of the FE model as additional unknowns in the Newton-Raphson

iterations. In his work, Nakata [34–36] developed a simplified inverse model for one-

at-a-time treatment of the dimensions of a group of permanent magnets to determine

the optimal shape and size required for producing prescribed flux densities at desired

points in the air-gap. As reported later, his developed method failed to produce

meaningful results for relatively more complicated design cases [37]. This was in

part due to the fact that the treatment of the boundary nodes as design variables

results in piecewise linear structures, which in turn oftentimes led to undesirable

spikes and edges in the structure of the optimized shapes [38–40], thus necessitating

a subsequent manual smoothing with compromised performance [38]. In the following

years, through a methodological analysis of different formulations for solving inverse

electromagnetic problems, Guamieri et al. demonstrated that inverse problems can,

in essence, be reduced to problems of numerical optimization [41].

In parallel with the formation of inverse field problems, some investigators

formulated the design process as a classic optimization problem with two main

segments: an “optimizer” to determine the direction of parameter adjustment; and

an “analyzer” to calculate the performance criterion [42]. Armstrong et al. [42]


7

introduced a rudimentary automated approach for the design of electromagnetic

components using 2-D magnetostatic integral field solutions in linear media. The

investigators concluded that several attempts of user intervention in the design process

is required to yield satisfactory results. With a similar perspective, Gitosusastro et al.

[43] distinguished the two main segments of a systematic optimization procedure for

the design of electromagnetic devices based on FE field computation methods. Also

reported in their early investigations was the sensitivity analysis, which revealed, by

way of direct differentiation of the FE matrices, how the performance characteristics of

the machine are affected by changes in the design variables. In [44], Koh et al. studied

magnetic pole optimization of a PM motor for the reduction of cogging torque. A

similar sensitivity analysis was devised to study the correlation of the design variables

with cogging torque. Park et al. fully elaborated this method of sensitivity analysis

of 2-D magnetostatic problems in [45]. Hoole et al. [46] applied the same concepts to

identify the magnetic systems through external measurements of the electromagnetic

field.

Due to the limitations of computational power, early systematic design

optimization methodologies tended to use deterministic, as opposed to stochastic,

strategies/optimizers in their search algorithms [47]. Some early references on the

application of deterministic methods in electromagnetic problems can be found in the

works of Saldanha et al. [48, 49]. The deterministic algorithms are characterized with

very fast convergence provided that the objective function is convex and differentiable.

However, they face the risk of premature convergence to local minima in a practical

optimization problem where such hypotheses cannot be made a priori. This


8

disadvantage of the deterministic approach led to popularity of stochastic algorithms

in the design of non-linear and complex electromechanical devices, especially with

the advent of more powerful computational systems. Stochastic methods are more

reliable in finding the global minimum of non-differentiable non-convex functions

because these methods do not rely on gradient calculations [50]. However, they

require a large number of function evaluations, typically in the range of thousands,

to locate global optimum. An early reference on comparative analysis of several

search algorithms with zero, first, and second order characteristics including the

Monte-Carlo iteration (MC), the steepest descent (SD), conjugate gradient, and

quasi-Newton is a publication by Gottvald [51]. Gottvald’s study was focused

on a large class of magnetostatic optimization problems which can be extended

to complex electromagnetic and electromechanical systems. He noticed that a

multitude of the factors involved in design of these complex systems precludes the

prescription of a fastest method, however, the zero-order methods such as MC are

more effective not in terms of time-efficiency, but in terms of reliability, numerical

stability, generality, and convergence to global minimum [52]. Gottvald also stressed

the challenges in solving computational electromagnetics optimization problems such

as the existence of local minima, and non-convex subdomains. In subsequent

collaborations, Gottvald and Preis rigorously compared deterministic and stochastic

optimization algorithms in computational electromagnetics design problems [53–55] to

unequivocally conclude that the studied stochastic methods are far superior in terms

of robustness and generality for complex applications, and that their convergence rates

can be comparable to those of deterministic methods. A combination of stochastic


9

and deterministic techniques for taking advantage of desirable characteristics of each

approach has also been suggested by some investigators, e.g. by Drago et al. in [56]

or by Simkin and Trowbridge in [57].

Russenshuck [58] categorized the optimization of PM machines into constrained

and unconstrained vector-optimization problems and comparatively studied the

convergence of several optimization algorithms for these two types of problems using

two different surface mounted PM machines as the benchmark designs. Russenshuck’s

work was also novel in other aspects such as the introduction of the concept of Pareto-

optimality for design of PM machines in a multi-objective optimization problem,

or the employment of global quantities (e.g. power density, torque, and magnet

weight) as opposed to local quantities (e.g. flux distribution) in defining the objective

functions. He pointed out the difficulties of incorporating higher order deterministic

algorithms in optimization problems based on numerical field calculations which are

subject to field discretization errors. This issue was further elaborated by Hoole et

al. in [59]. They found out that the discontinuities in the objective function have

no physical basis, yet they will persist regardless of the mesh accuracy. In a later

work, the same investigators developed a structural mapping method for geometric

parametrization to minimize these discontinuities [60], which was then used in the

design optimization of salient pole synchronous machines [61].

In the early 1990s, several investigators reported the application of stochastic

search algorithms in the design optimization of PM machines based on numerical

field calculations. Preis et al. [62] utilized evolutionary algorithms for the design

optimization of the pole shape of a dipole magnets using two- and three-dimensional
10

FE models. Some investigators, e.g. Schafer Jotter and Muller in [63] or Simkin

and Trowbridge [64] reported a detailed step-by-step procedure for optimization of

electromechanical devices using numerical techniques and systematic optimization

algorithms.

1.2.2 Design Optimization for Rated Operation

Owing to the developments in electromagnetic FEM and the increase of processing

power of computational resources, population-based design optimization of electric

machines using high fidelity FE models has become an established practice. In

[65], Uler et al. developed a population-based method for design optimization

of electromagnetic devices using 2-D FE models as the analysis tool and genetic

algorithms (GA) as the search algorithm. The investigators implemented the method

in Fig. 1.3 for design optimization of the pole face of a motor to achieve sinusoidal

magnetic flux density in the air-gap for one particular loading condition, i.e. a

current density of 10A/mm2 . Four design parameters each coded into 16-bit string

segments of binary codes were incorporated in a large-scale design optimization with a

population size of 30 members which was carried out over 24 iterations. In subsequent

research, Mohammed [66] discussed the practical issues in the application of GA to

design optimization of electromagnetics devices, with a particular emphasis on the

issue of premature convergence.

In [67], Chun et al. developed an evolutionary strategy by combining genetic

algorithm and simulated annealing for minimization of cogging torque of a closed

slot PM servo motor. The FEM was used for performance evaluation of the design
11

Start design
optimization

● Randomly created initial population

● Fitness of the design candidate


● FEA
● Average fitness of population

Avg. fitness Stop

● Reproduction, crossover, and mutation


● New generation

Figure 1.3: Design optimization based on GA.

candidates. A rated torque of 3 Nm and an ampere turn of 136 AT was considered.

The optimization was performed over 900 generations each of 1 member to optimize

the structure of the stator slot which was parameterized through 31 design variables.

Also in [68], Chung et al. optimized the slot shape of a PM machine to reduce the

cogging torque using FEM and a similar evolutionary algorithm. Their parametrized

model consisted of 3 design variables.

In [69], Bianchi and Bolognani utilized GA to optimize a surface mounted PM

motor using both analytical and FE models for analysis of the motor performance.

The method could take global figures of merit such as torque, efficiency, and the

material cost into account. Using the FE-based model, the investigators pursued

two scenarios of optimization with different objectives, one with the objective of

minimizing the PM mass for a rated torque of 10 Nm, and the other with the objective

of maximizing torque density within confined stator outer diameter and axial stack
12

length. In both cases, the model consisted of five design parameters on the stator

and PM structures. Furthermore, additional performance constraints on the winding

temperature and the minimum air-gap flux density were defined. The optimization

was performed with different population sizes to conclude that the best results are

achieved with a population size of larger than 50 members. The investigators also

performed the optimization using hill-climbing direct search method to observe that

the optimized designs obtained using GA were slightly better than those produced

by the hill-climbing direct-search method, indicating the presence of local minima in

the search space. In a similar work, Bianchi and Canova [70] used the same GA-

FEM large-scale design optimization methodologies to maximize the torque of an

IPM machine for a given peak stator current of 70 A. In this case, the optimal values

of five design parameters were pursued in a procedure consisting of 10 generations

each of 100 members.

In early applications of FEM in the large-scale design optimization of PM

machines, due to the computationally demanding nature of the FE models, typically

a single magnetostatic solution would be performed to characterize the machine

performance. In [71], Ohnishi and Takahashi included the rotor motion by taking

into account the relative position of the rotor with respect to the stator through

a nonconforming meshing technique, first introduced by Fouad et al. [72], which

reduces the computation time required for mesh modification during the rotation.

The investigators then used this technique in optimizing the rotor of an IPM machine

with four design parameters. The optimization which was carried out over a large

number of iterations, led to significant reductions of PM volume and torque ripple


13

under a rated current of 3 A(rms). Many other investigators have aimed to reduce

the computational burden of the design optimization procedure by incorporating

alternative modeling techniques, including analytical models as in [73], lumped

parameter models as in [74], and artificial intelligence models as in [75]. However, as

studied in [76–78] for a broad range of modeling techniques, there is always a trade-off

between accuracy and execution time of the model-based design and analysis.

In the case of sinusoidally excited three-phase PM synchronous machines, as

was observed by some investigators in [72, 79–81], in addition to the magnetic

periodicity, the existing electrical symmetry enables estimation of various performance

metrics of the machine by only analyzing a sixth of an electrical period. This

idea was fully elaborated by Sizov et al. in [82] as the computationally efficient-

finite element analysis (CE-FEA). The CE-FEA approach was subsequently utilized

in the parametric study and later-on in large-scale design optimizations of several

PM machines using only a regular desktop workstation, as opposed to using high

performance computing (HPC) systems as reported in [83]. Although the CE-

FEA method does not provide an estimation of the rotor core losses, these losses

are usually smaller than the stator core losses due to the unvarying field in the

stator of synchronous machines. In [84], five case-study machines with different

pole-slot combinations and power ratings were optimized using CE-FEA and DE.

Each case-study was fully parameterized to allow variations of the rotor and

stator structures. The optimization involved thousands of design evaluations each

incorporating multiple magnetostatic solutions to take into account the rotor motion.

The designs were optimized to maximize the torque density, and to minimize the
14

torque ripple and power losses at the rated load operating point. In [85], Zhang

et al. reported the optimization of a fault-tolerant 12-slot 10-pole IPM machine

using the CE-FEA and DE design synthesis package. The motor model, which was

parameterized through 9 independent design variables, was optimized for minimized

losses and active material cost, subject to several performance constraints on torque

ripple, induced voltage THD, and degree of PM demagnetization, all at the rated

load point corresponding to a torque of 42 Nm and a speed of 1800 r/min. For

a large number of design candidates, i.e. 50 generations of 70 members each, the

optimization took only 28 hours on a PC workstation. The same design methodology

was employed by Brown et al. [86], for optimization, comparison, and prototyping of

two IPM topologies for use in a commercial heat pump. The investigators concluded

that the CE-FEA optimization methodology was successfully tested in designing the

high-efficiency prototypes. Duan et al. in [87], Zhang et al. in [85, 88], and Wang et

al. in [89, 90] reported several other successful implementations of the CE-FEA design

optimization methodology for improving the rated performance of various types of

PM or SyR machines.

1.2.3 Design Optimization for Entire Operating Range

One of the major merits of salient pole synchronous machines is their efficient

operation in the extended speed range. This is due to the contribution of the

reluctance torque in this mode of operation, which can be overlooked if not directly

evaluated in the optimization fitness function. However, the common practice is to

assess the fitness function at the rated load point.


15

In [91], Sibande et al. implemented an FE-based design optimization algorithm

to improve the field weakening performance of a synchronous reluctance machine

(SyRM) which was equipped with additional sheets of PM in between the rotor

flux barriers resulting in a PM-assisted SyRM. The goal of the optimization was

to minimize the volume of the rotor PMs while realizing the performance constraints

on minimum required torque and maximum allowed supply voltage under rated and

maximum speeds. The design parameters were limited to the thickness of the magnet

sheets in the rotor flux barriers, leaving the stator and the rotor structures unchanged.

An optimal design was achieved and was shown to outperfom the original SyRM

in terms of torque and terminal voltage characteristics. The efficiency during the

extended speed operation was not investigated.

Zarko et al. first introduced a systematic optimization algorithm for

simultaneously minimizing the rotor volume and maximizing the power capability

in the flux weakening regime in [92] based on the idea of equality of the rated

current and the characteristic current [93], i.e. the ratio of magnet flux linkage to

d-axis inductance. The investigators utilized FE-based models and a DE optimizer

for maximizing the characteristic current and minimizing the active volume of the

example two- and three-layer IPM motors. An additional performance constraint

was imposed on the back emf to limit it to 400 V at maximum 6 000 r/min speed.

A relatively large number of design parameters, 13 and 15 variables depending on

the number of the magnet layers, and a relatively large number of design candidates,

8 000 and 10 000 designs, were evaluated. Although a minimum efficiency of 0.8 was

considered as a design constraint, maximizing efficiency was not pursued. Another


16

pioneering model-based design optimization study for improving the field weakening

performance of IPM machines was developed by Quyang et al. [94], where Monte

Carlo and DE algorithms were coupled to FE-based models to optimize two IPM

motors with modular and regular stator structures. In the optimization of the

conventional stator design, similar to Zarko’s work in [92], the motor was optimized

for maximum electromagnetic torque at base speed, and maximum normalized

characteristic current for improved field weakening operation.

Similarly, the idea of equality of the characteristic current, ICH , with the rated

current, IR , was introduced as an additional objective for optimization of IPM motors

for enhanced field weakening performance. In [95], Pellegrino and Cupertino adopted

this criterion to optimize the rotor structure of an IPM motor with three PM layers

using a genetic algorithm optimizer coupled to a static FE-based model. Through

performing four sets of optimizations with different objectives, it was asserted

that there is a trade-off between the constant power range and torque production

capability. In [96], Zhang et al. developed an optimization algorithm based on CE-

FEA coupled to a DE optimizer to minimize the difference between the values of ICH

and IR , in order to achieve a wide constant power range of operation. Minimization of

material cost and stator losses at the rated load point were simultaneously pursued.

Based on these objectives, it was concluded that for the two studied IPM motors with

single-layer and double-layer PMs, when extended speed operation is considered, high

efficiency, high power factor, and saliency increase the material cost.

The equality of ICH , and IR improves the torque production capability of the

machine [93]. However, from the efficiency standpoint, congruity of ICH , and IR
17

cannot be the ideal criterion for constant power operation. By definition, maximizing

ICH results in either stronger magnets or lower d-axis inductance. Accordingly, when

the non-linear and lossy nature of the machine is considered, for machines with larger

ICH , a larger magnitude of negative d-axis current is required to demagnetize the

magnet flux and maintain controllability in the extended speed range. This results

in higher copper losses and diminishes the energy efficiency of the field weakening

regimen [97, 98].

In some studies, the field weakening performance of the machine is directly

investigated by evaluating two critical load points: maximum torque at base and

maximum speeds. In [99], Parasiliti used an FE-based model with a controlled

random search optimizer to realize three objectives: maximize the torque at base

speed, maximize the torque at maximum speed, and minimize the weight of the motor.

Although the equality of ICH , and IR is not directly introduced in the optimization

fitness function, overlooking efficiency in the definition of the objectives resulted in

an optimized design in which the two currents are equal. Alternatively, in [100] in

lieu of minimizing the weight, maximizing the torque and efficiency at the base and

maximum speeds were considered as the objectives by Lee and Kwon. In this reference

[100], an FE-based model in conjunction with a so-called Kriging approximation

method were used to simultaneously improve the efficiency and operation range of

a concentrated flux IPM motor. One limitation of the proposed method in [100], is

the estimation of the stator winding excitation current using inductance-based motor

models. Also in [101], Yamazaki et al. optimized the rotor design of an IPM motor

through FE-based models to minimize the iron losses for maximum torque operation
18

at base and maximum speeds. However, the rotor geometry was only subjected to

limited variations with some additional assumptions regarding the current advance

angle which were not fully substantiated. The same investigators carried out another

fine-tuning optimization of the rotor shape of IPM motors for reducing the iron losses

in the field weakening region in [102].

In a different approach [103], Cupertino et al. suggested a three-stage FE-based

optimization with genetic algorithm optimizer, for achieving desirable field weakening

performance in a PM assisted SyR motor. In the first stage, the SyR motor with

three flux barriers is optimized for maximum torque, and minimum torque ripple at

the rated load point. Upon finding an optimal solution, a second optimization is

carried out with constrained design parameters close to those obtained in the first

stage plus or minus 15%. In the final stage of the design, PMs are introduced in the

rotor structure. The grades of the PMs are so chosen to achieve the required flux

weakening operation. Another study on PM assisted SyR motor by Barcaro et al. in

[104] analyzed the influence of the PM volume on the field weakening performance

using FEM. Subsequently, the relationships between PM remnant flux density and

efficiency at two load points residing at base and corner speeds were investigated.

In more recent studies, researchers have attempted to optimize the machine

performance over the entire driving cycle. In [105, 106], Wang et al. proposed a

method known as the cyclic representative points to efficiently model a target driving

cycle, specifically the New European Driving Cycle (NEDC) and the Artemis Urban

Driving Cycle, by a finite number of torque versus speed points. These points were

derived based on the energy distribution function specifically calculated for the vehicle
19

model and the target driving cycle. The investigators subsequently performed the

optimization over these cyclic representative points. However, these studies do not

fully address the following two issues:

1. The selection criterion for the representative points is relatively subjective,

requiring a systematic approach for identification of the high-energy-throughput

zones of the machine energy distribution function.

2. As opposed to a large-scale optimization algorithm, the method undertaken in

those studies is a comparative study of a few designs with limited number of

design variables.

In [107] the second shortcoming of [105, 106] was addressed in optimizing a PM

assisted SyR motor using a FE-based model with genetic algorithm optimizer by

Carraro et al. However, only two representative points, which were selected in a

relatively subjective manner, were considered. This limited number of points cannot

effectively represent the driving cycle. In the case study, both of the two representative

points resided in the constant torque range. Therefore, the objectives which were

defined as minimization of the torque ripple and motor losses over these points do

not necessarily hold through the entire operating range.

1.3 Statement of the Problem

A typical traditional design procedure of an electric machine is shown in Figure

1.4. When an initial design is obtained, further improvements are possible through

numerical optimization.
20

Requirements Decisions Validation

 Torque profile  Suitable motor type  Electromagnetic


 Speed profile  Number of phases performance
 Losses and efficiency  Pole-slot combination  Thermal performance
 Size/Weight  Sizing  Stress analysis/
 Cogging/Ripple  Stator/Slot structure mechanical
 Terminal conditions  Rotor layout performance
(EMFs, PF, R, L, etc.)  Winding configuration  Transient performance
 Noise/Vibration  Type of drive  …
 Manufacturability  Losses and cooling
 Cost  …
 …

Figure 1.4: High level design flow of an electric machine.

In the light of the literature review, the common practice in large-scale design

optimization of electric machines is to improve the desired performance metrics

under the rated operating conditions. This is most suitable for applications in which

the load profile is constant over the majority of the operation period, e.g. classes

S1-S3, and S6 of standard rotating electrical machines according to IEC 60034-1.

However, in many applications, the motor torque and speed profiles experience wide

variations, e.g. electric motors in traction applications, or classes S4, S5, and S7-S10

of standard electric machines according to IEC 60034-1. Many impediments exist in

taking the entire range of operation into account, of which the most notable is the

computationally demanding nature of various stages of the optimization procedure.

Accordingly, this dissertation addresses the challenges in the design optimization

of high performance electric machines with non-constant operating cycles where an

optimal design is to maintain high performance under wide loading conditions.


21

1.4 Dissertation Organization

In view of the problem background and the literature search, Chapter 2 discusses

the essence of a recently introduced computationally efficient approach for large-

scale design optimization of electric machines under rated operating conditions. This

automated design package, which serves as the starting point of the topics investigated

in this dissertation, is utilized in Chapter 2 to optimize the performance of the

well-established benchmark design represented by Toyota Prius Gen 2 IPM motor

for one particular operating condition. The necessity of including the entire range

of operation into the optimization procedure is highlighted by this example. To

mathematically corroborate the importance of design optimization over the motor

operating cycle, the variations of optimal design rules of PM machines with respect

to ampere loading level are systematically investigated in Chapter 3. For this

purpose, two IPM motors with different cooling systems and winding configurations

are studied. As the first step for increasing the computational efficiency of the

design optimization procedure, a fast stochastic search algorithm for large-scale

design optimization of electric machines is introduced in Chapter 4. By reducing

the computational burden of the searching process, the developed algorithm enables

adopting more complicated models with larger number of load operating points. In

Chapter 5, the drive-cycle design optimization procedure, consisting of several stages,

is developed, followed by optimization of two benchmark traction motors in Chapter 6.

Chapter 7 is dedicated to additional implications of design of PM machines, including

a sensitivity analysis of active material cost with respect to commodity price changes,
22

a computationally efficient method for calculation of strand eddy current losses, and

a method for estimation of tangential mechanical stresses on the rotor bridges. The

conclusions, and future works are provided in Chapter 8.

1.5 Related Publications

So far the topics discussed in this dissertation are published, or are in press in various

IEEE publications, as follows:

[108] A. Fatemi; D. M. Ionel; N. A. O. Demerdash; T. W. Nehl, “Optimal Design

of IPM Motors with Different Cooling Systems and Winding Configurations,”

in IEEE Transactions on Industry Applications , vol.PP, no.99, pp.1-1 (Early

Access)

[109] A. Fatemi; N. Demerdash; T. Nehl; D. Ionel, “Large-scale Design Optimization

of PM Machines Over a Target Operating Cycle,” in IEEE Transactions on

Industry Applications , vol.PP, no.99, pp.1-1 (Early Access)

[110] A. Fatemi; D. M. Ionel; N. A. O. Demerdash; T. W. Nehl, “Fast Multi-

Objective CMODE-Type Optimization of PM Machines Using Multicore

Desktop Computers,” in IEEE Transactions on Industry Applications , vol.PP,

no.99, pp.1-1 (Early Access)

[111] A. Fatemi, D. M. Ionel, M. Popescu, and N. A. O. Demerdash, “Design

Optimization of Spoke-Type PM Motors for Formula E Racing Cars,”

Accepted for 2016 IEEE Energy Conversion Congress and Exposition (ECCE),
23

Milwaukee, WI, 2016.

[112] A. Fatemi, D. M. Ionel, N. A. O. Demerdash, S. Stretz, and T. M. Jahns

“RSM-DE-ANN Method for Sensitivity Analysis of Active Material Cost in PM

Motors,” Accepted for 2016 IEEE Energy Conversion Congress and Exposition

(ECCE), Milwaukee, WI, 2016.

[113] A. Fatemi, D. M. Ionel, N. A. O. Demerdash, D. Staton, R. Wrobel, and Y.

C. Chong, “Computationally Efficient Method for Calculation of Strand Eddy

Current Losses in Stator Windings of Electric Machines,” Accepted for 2016

IEEE Energy Conversion Congress and Exposition (ECCE), Milwaukee, WI,

2016.

[114] A. Fatemi, N. A. O. Demerdash, D. M. Ionel and T. W. Nehl, “Large-scale

Electromagnetic Design Optimization of PM Machines Over a Target Operating

Cycle,” 2015 IEEE Energy Conversion Congress and Exposition (ECCE),

Montreal, QC, 2015, pp. 4383-4390.

[115] A. Fatemi, D. M. Ionel, N. A. O. Demerdash and T. W. Nehl, “Fast

Multi-Objective CMODE-type Optimization of Electric Machines for Multicore

Desktop Computers,” 2015 IEEE Energy Conversion Congress and Exposition

(ECCE), Montreal, QC, 2015, pp. 5593-5600.

[116] A. Fatemi, D. M. Ionel and N. A. O. Demerdash, “Identification of Design

Rules for Interior PM Motors with Different Cooling Systems,” 2015 IEEE

International Electric Machines & Drives Conference (IEMDC), Coeur d’Alene,


24

ID, 2015, pp. 1228-1234.

[98] A. Fatemi, N. A. O. Demerdash and D. M. Ionel, “Design Optimization of IPM

Machines for Efficient Operation in Extended Speed Range,” Transportation

Electrification Conference and Expo (ITEC), 2015 IEEE, Dearborn, MI, 2015,

pp. 1-8.
25

CHAPTER 2
DESIGN OPTIMIZATION FOR RATED
OPERATING CONDITIONS

In this chapter, an automated design package for optimization of sine-wave driven

permanent magnet synchronous machines (PMSMs) and synchronous reluctance

machines (SyRM) is reviewed. The presented design optimization methodology was

previously developed by Dr. Gennadi Sizov [117] and Dr. Peng Zhang [118] at the

Electric Machines and Drives Laboratory at Marquette University over the past few

years. This design package, which on par with the common practice was developed

for improving the machine rated performance, serves as the starting point of the

research conducted in this dissertation. First, the essence of this design optimization

approach including its two main segments, one for computationally efficient FE-based

performance evaluation of the design members, and another for fast convergence to

the globally optimal design solutions, is reviewed. Subsequently, a typical traction

propulsion motor represented by the well-established/well-documented Toyota Prius

Gen 2 interior PMSM design will be optimized for its peak operating condition.

Through this case study, the principles of the automated design package developed

by the team at Marquette University are reviewed. Furthermore, the shortcomings

of the common optimization techniques in overlooking the performance requirements

throughout the entire range of operation are delineated.


26

2.1 Design Synthesis Using CE-FEA and DE

The design synthesis of electric machines is a classical optimization problem which

aims at realizing a set of objectives under a set of constraints. Either global

performance metrics such as torque, losses and efficiency, terminal conditions, and

material cost, or local quantities such as air-gap flux distribution, and PM local

demagnetization can constitute the set of objectives. The constraints can also be

either geometry-related, such as those imposed by mechanical design considerations,

or performance-related, such as maximum permissible torque ripple, or degree of PM

demagnetization.

A recently developed large-scale design optimization approach [119], which

features FE-based performance evaluation of thousands of design members within

a reduced time, is composed of two main parts, see Fig. 2.1. The first part is a

computationally efficient-finite element analysis (CE-FEA) [80] which is utilized for

performance evaluation of the design members with high fidelity. The second part is

a stochastic differential evolution (DE) [120] search algorithm which is employed for

finding the globally optimal solutions. In this section, the principles of CE-FEA and

DE are reviewed.

2.1.1 Overview of Computationally Efficient-Finite Element

Analysis

The CE-FEA approach fully exploits the existing electric symmetry and magnetic

periodicity of PMSMs with sinusoidal current excitation. It is based on a sequence


27

Differential Evolution Computationally Efficient -FEA

● Parameterized 2D model
Params. F and Cr
● Magnetostatic analysis settings

Estimate design Max. Torque Per Amp.


Generate new population of Np Optimum torque angle
candidate designs: performance metrics:
● Design from DE
● Mutation and Crossover Check ● Torque(em, shaft,
cogging, ripple) Flux-weakening operation
● Check boundaries of design params. Optimization at single or
● Motor performances from CE-FEA ● Losses (iron, copper, multi-operating points
● Design objectives and constraints PM)
● Iterate until convergence criteria are ● Active material costs
satisfied. ● PM demagnetization Parallel Execution
● Estimate the end effects On Multiple CPUs

Figure 2.1: Design optimization of PM machines based on CE-FEA and DE.

of two-dimensional (2D) magnetostatic FE solutions at successive rotor positions

covering one sixth of an electrical revolution/cycle. Compared to full-fledged transient

time-stepping FEA, analyzing the motor performance using the CE-FEA approach

can lead to significant savings in the modeling time, up to two orders of magnitude

as stated in [119]. The principles of this modeling approach are briefly explained as

given next.

2.1.1.1 Reconstruction of Field Waveforms

The steady-state 2D magnetostatic FEA of a PM machine in xy-plane is expressed

by the Poison’s equation given below:


   
∂ ∂A ∂ ∂A
ν + ν = −J − JP M (2.1.1)
∂x ∂x ∂y ∂y

where ν is the material reluctivity, A is the magnetic vector potential (mvp), J is the

excitation current density, and JP M is the equivalent current density for modeling
28

e'2
e''2 e2

b1+ c2- c1-


b2+ a2+
a2- e'1 a1+
e''1
e1

(a) 48-slot 8-pole configuration

e'2

b2+ b1+ e'1


b2- b1-
c1+ a2+

e''1

e''2 e2
c1- a2-

c2- a1-

e1

c2+ a1+

(b) 12-slot 10-pole configuration

Figure 2.2: Mapping the field values between typical sister elements in the CE-FEA
method.

the permanent magnets. Equation (2.1.1) is solved at successive rotor positions for

obtaining the unknown mvps according to a predefined excitation, J and JP M . The

instantaneous values of J are determined from the rotor position, θm , and loading

condition of the machine.

If a balanced three-phase sinusoidal current excitation is assumed, the electric

symmetry in the stator geometry can be used to map the tangential, fT , and radial,

fR , field values between sister elements throughout the stator periodicity span, as
29

2
Reconstructed Btooth

Mid−tooth B (T)
1 Samples tooth−a
Samples tooth−b
0 Samples tooth−c

−1

−2
0 60 120 180 240 300 360
Rotor position (deg. el.)

(a) 48-slot 8-pole configuration

1.5
Reconstructed Btooth
1
Mid−tooth B (T)

Samples tooth−a
0.5 Samples tooth−b
0 Samples tooth−c

−0.5
−1
−1.5
0 60 120 180 240 300 360
Rotor position (deg. el.)

(b) 12-slot 10-pole configuration

Figure 2.3: Reconstruction of mid-tooth flux densities using the CE-FEA method.

given below:
kαs P
fR,T (t + , r, θ) = fR,T (t, r, θ + kαs ) (2.1.2)

where, k is the index that depends on the slot-pole combination and winding layout,

P is the number of poles, ω is angular frequency, and αs is the slot pitch in mechanical

measure.

Typical sister elements for two different pole-slot combinations, i.e. a 48-slot 8-

pole, and a 12-slot 10-pole configuration are graphically illustrated in Fig. 2.2 (a)

and (b). As shown Fig. 2.2, in the CE-FEA method, the field values are mapped
30

1.5
Reconstructed Byoke
1

Mid−yoke B (T)
Samples yoke−a
0.5 Samples yoke−b
0 Samples yoke−c

−0.5
−1
−1.5
0 60 120 180 240 300 360
Rotor position (deg. el.)

(a) 48-slot 8-pole configuration

1.5
Reconstructed Byoke
1
Mid−yoke B (T)

Samples yoke−a
0.5 Samples yoke−b
0 Samples yoke−c

−0.5
−1
−1.5
0 60 120 180 240 300 360
Rotor position (deg. el.)

(b) 12-slot 10-pole configuration

Figure 2.4: Reconstruction of mid-yoke flux densities using the CE-FEA method.

between typical sister elements such as e1 , e01 , and e001 or e2 , e02 , and e002 . It should

be pointed out that the relative location of the sister elements is independent of the

rotor PM layout as long as the number of the poles is fixed.

The aforementioned existing symmetry in the stator structure of sinusoidally

excited PMSMs is fully utilized in the CE-FEA method to reconstruct the entire

field waveforms through multiple snapshots of magnetostatic FE solutions over a

time span corresponding to 60 electrical degrees.

In Fig. 2.3, the reconstruction of the mid-tooth flux densities using CE-FEA
31

2
Reconstructed λ
a
1 Samples phase−a
Samples phase−b

λ (Wb)
0 Samples phase−c

a −1

−2
0 60 120 180 240 300 360
Rotor position (deg. el.)

(a) 48-slot 8-pole configuration

1
Reconstructed λ
a
0.5 Samples phase−a
Samples phase−b
λa (Wb)

0 Samples phase−c

−0.5

−1
0 60 120 180 240 300 360
Rotor position (deg. el.)

(b) 12-slot 10-pole configuration

Figure 2.5: Reconstruction of stator winding flux linkages using the CE-FEA method.

approach are illustrated for the two example machines previously shown in Fig. 2.2,

respectively using 10 and 7 static solutions for the 48-slot 8-pole, and 12-slot 10-pole

machine configurations.

As shown in Fig. 2.4 for the same machine configurations, a similar procedure

can be employed for reconstruction of the mid-yoke flux density waveforms. It should

be pointed out that for illustration purposes in Figs. 2.3 and 2.4, the “virtual search

coils” [118] for collecting the flux density values are located perpendicular to the main

flux path. Otherwise, the tangential and radial components of elemental flux densities
32

40
Reconstructed torque
35 Samples

Torque (Nm)
30

25

20
0 60 120 180 240 300 360
Rotor position (deg. el.)

(a) 48-slot 8-pole configuration

32
Reconstructed torque
31 Samples
Torque (Nm)

30

29

28
0 60 120 180 240 300 360
Rotor position (deg. el.)

(b) 12-slot 10-pole configuration

Figure 2.6: Reconstruction of the torque profile using the CE-FEA method.

should be treated separately.

In addition to the elemental flux density values in the stator core, the same

symmetry rules can be found in the waveforms of the stator winding flux linkages.

This is illustrated in Fig. 2.5(a) and (b) for the two previously discussed example

machines. The waveforms of the induced voltages can be subsequently obtained from

the differentiation of the flux linkage waveforms.


33

2.1.1.2 Reconstruction of Torque

In the CE-FEA method, the torque values, which can be calculated using the virtual

work method [121] or the Maxwell stress tensor, are obtained at the sample points

within the first 60 electrical degrees of the fundamental cycle. Subsequently, these

sample points are replicated over the entire cycle to reconstruct the torque profile.

A Fourier analysis on the resultant torque profile is then conducted to extract the

average torque in Eq. (2.1.3) and the torque ripple in Eq. (2.1.4) [118]:

X
Tem = Tavg + Tn cos (nωt + φn ) (2.1.3)
n=6,12

Tr = (Tmax − Tmin )/Tave × 100 (2.1.4)

The process of reconstruction of the torque profile in the CE-FEA method is

illustrated in Fig. 2.6 for the two previously discussed 48-slot 8-pole, and 12-slot

10-pole machines.

2.1.1.3 Estimation of Core Losses

Upon deriving the waveform profiles of the field quantities, a subsequent Fourier

analysis should be performed according to Eq. 2.1.5 to determine their harmonics

expressions:
nX
max

f (θ) = Fn cos(nθ + φn ) (2.1.5)


n=1

where Fn and φn are the magnitude and the phase angle of the nth harmonic,

respectively.

The Fourier series of the stator core flux densities is of special interest for

calculation of the hysteresis losses, Ph , and eddy-current losses, Pe , which constitute


34

the total core losses, PF e , of the stator laminations. Throughout this work, the CAL2

loss model introduced by Ionel et al. in [122], which features accurate estimation of

core loss components at individual frequencies, is used:

PF e = kh (B, ω)ωB 2 + ke (B, ω)(ωB)2 (2.1.6)

According to the loss model in Eq. (2.1.6), the hysteresis, kh , and eddy-current,

ke , core loss coefficients, are expressed as functions of the peak flux density, B, and

angular frequency, ω, i.e. kh (B, ω), and ke (B, ω). It has been previously demonstrated

[123–125] that the frequency dependency in these functions can be considered at

multiple non-overlapping frequency ranges, thus simplifying the expressions for core

loss coefficients, as given in Eq. (2.1.7).



kh (B) = kh3 B 3 + kh2 B 2 + kh1 B + kh0
(2.1.7)
k (B) = k B 3 + k B 2 + k B + k
e e3 e2 e1 e0

Accordingly, the hysteresis and eddy current loss components can be separately

expressed based on Fourier series derivations of the stator core flux densities using

the CE-FEA method:

nX
max
ω
Ph = kh (Bn )(n )B 2 , (W/kg) (2.1.8)
n=1
2π n

nX
max
ω 2 2
Pe = kc (Bn )(n ) Bn , (W/kg) (2.1.9)
n=1

The CE-FEA method does not provide an indication of the rotor losses, i.e. core

and magnet losses. Nonetheless, these losses are expected to be smaller, as opposed

to the stator core losses, due to the non-varying (non-pulsating) fundamental field in

the rotor laminations. Whenever high frequency fluctuations (pulsations) of the rotor
35

field are significant, which is more likely to occur in PM machines with fractional

slot concentrated winding configurations [126], the rotor losses can be taken into

account by using alternative surrogate methods such as response surface methodology

[80, 119].

2.1.1.4 Estimation of Ohmic Losses

The dc copper losses, Pdc , are obtained by Eq. (2.1.10) provided below:

Pdc = vCu ρ · J 2 , (W ) (2.1.10)

where vCu is the volume of the copper including the end-turns, J is the excitation

current density, and ρ is the copper resistivity which is calculated for a given

temperature, T emp, by using a reference temperature of 20◦ C in Eq. (2.1.11).

ρ = 1.724 × 10(−8) [1 + 0.00393(T emp − 20)], (Ω · m) (2.1.11)

2.1.2 Overview of Differential Evolution

Finding the globally optimal design solutions requires incorporation of a search

algorithm, aka optimizer. The optimizer utilized in the design synthesis package

developed by the team at Marquette University is the differential evolution (DE)

algorithm [120].

Similar to other evolutionary algorithms, the DE algorithm is a stochastic

optimizer. That is, there is a degree of uncertainty involved in the evolution model

of the “non-dominated” designs. Yet, the DE algorithm is characterized by its

unique method for the generation of trial design members, ~u, which are the designs
36

that compete with the parent population members, ~xg , to determine the offspring

population, ~xg+1 . The DE algorithm with adaptations for electric machinery design

optimization problems is explained in the following discussion.

2.1.2.1 Initialization

In a typical 2-D design problem, the cross section of the given machine configuration

can be parametrized by defining a certain number, p, of design parameters establishing

the strongest correlations with the machine performance metrics. Accordingly, each

design member, ~x, is a vector consisting of p design parameters, ~x = [x1 , x2 , ..., xp ].

Each generation of the DE optimization algorithm is composed of Np distinct

designs which constitute the current population, P~g = [~xg,1 , ~xg,2 , ..., ~xg,Np ]. The first

generation of the design members is randomly generated by considering the predefined

upper, xi,max , and lower, xi,min , bounds of the design parameters as given below:

xi = xi,min + rand(0, 1)(xi,max − xi,min ) (2.1.12)

The number of the design parameters p is determined by the machine model,

whereas the population size, Np , is determined heuristically to yield the fastest

rate of convergence to the globally optimum solution. The population size, Np , is

recommended to be at least 5-10 times greater than p for optimization of PM machines

using the classical DE approach [117, 118].

2.1.2.2 Generation of Trial Designs

Upon identification of the design members of the parent population, and calculation

of their performance metrics, the offspring population needs to be determined by


37

comparing the performance of the design members in the parent population with

that of a corresponding trial design. Each parameter of the trial design, ~uk , is created

in reference to a corresponding design member, ~xk , in addition to three randomly

selected design members from the current population, ~xr0 , ~xr1 , and ~xr2 , through the

mutation and crossover operations expressed below:



xr0 + F (xr1 − xr2 ) if rand(0, 1) ≤ Cr ,
ui = (2.1.13)
x otherwise
i

where F is a positive scale factor, F > 0 , and Cr is the crossover probability,

Cr ∈ [0, 1], i.e. 0 ≤ Cr ≤ 1.0. Although similar to Np , the specific values of F and Cr

are not known, it is recommended that F < 1 for better reliability and convergence

rate [120]. Furthermore, a lower limit of Fmin for enhanced diversity of the population

is suggested in [120] as given in Eq. 2.1.14 below:


s
(1 − Cr /2)
Fmin = (2.1.14)
Np

If the lower or upper bounds of the design parameters are violated in the trial

member, either Eq. (2.1.12) or the “bounce-back” method in Eq. (2.1.15) given

below can be used to recreate the trial member.



xi + rand(0, 1)(xi,min − xi ) if ui < xi,min
ui = (2.1.15)
i,max − xi ) if ui > xi,max
x + rand(0, 1)(x
i

2.1.2.3 Selection

After the performance metrics of the trial designs are identified, in this case through

utilization of CE-FEA, a competition takes place between corresponding pairs of

members to determine the winning offspring design. This competition is based on


38

x1,g x2,g ... xNp,g


...
rand (0,1) < Cr → ui = xr 0 + F ( xr1 − xr 2 ) xr1,g
else xr2,g ...
ui = xi xr3,g
...
u1,g u2,g ... uNp,g

CE-FEA ...

x1,g VS u1,g ...

x1,g+1 x2,g+1 ... xNp,g+1

Figure 2.7: Evolution model of DE from one generation to another. Multiple


members of the current population can be processed simultaneously to decrease the
computation time.

minimizing/maximizing a set of objectives, and satisfying a set of equality/inequality

constraints defined in the optimization fitness function. In a typical electric machinery

design problem, the minimization of active material cost and power losses are

common objectives, subject to performance constraints on maximum torque ripple or

maximum degree of PM demagnetization.

According to the Lampinen method of constraint handling which is specifically

recommended for DE algorithms [120], the trial vector ~u wins the competition if:

• it has better or equal objectives and satisfies all the constraints, or

• it does not violate the constraints whereas the current design does, or

• its constraint violation is less severe than the current member.


39

In the case of a multiobjective design optimization, the Pareto optimality can be

used to determine the non-dominated design members. According to the Pareto

criterion, in a minimization problem, a design is non-dominated if none of its

objectives are higher and at least one of its objectives is lower compared to the

other design members.

The evolution model of DE coupled to CE-FEA for two consecutive generations

is illustrated in Fig. 2.7.

2.1.2.4 Termination

Although there is not a specific rule for termination of the optimization iterations, a

few possible stop criteria for the DE algorithm are discussed in [77, 127, 128]. The

general idea is that the optimization needs to be carried out until the changes in the

value of the objective functions between two consecutive generations are small. From

this point on, the optimization can be continued to add to the density or diversity of

the optimal designs in the optimal solution region, which is in the form of a Pareto

front in case the number of objectives is greater than one.

2.2 Case Study Design Optimization of Toyota

Prius Gen 2 Motor at Its Nominal Operating

Point

In this section, the 48-slot 8-pole IPM machine topology shown in Fig. 2.8 which

was used in the Toyota Prius Gen 2 Hybrid Synergy Drive (HSD), is optimized at
40

Figure 2.8: Cross-section of the Toyota Prius Gen 2 IPM motor.

the nominal operating point using the design optimization package described in the

previous section. The purpose of this case study design is twofold: (a) to examine the

steps involved in a sample case study using CE-FEA and DE, and (b) to illustrate

the shortcomings of the existing approach in performance improvement for the entire

range of operation.

2.2.1 Parametric CE-FEA Model

The details concerning the general nature of a finite element non-linear magnetic

field computation, such as setting up the geometries, winding parameters, material

types, operating conditions, meshing, boundary conditions, etc., are available in

the literature [21–31] and, therefore, will not be discussed here. In addition to

the fundamental procedures involved in solving a magnetostatic solution, which in


41

this case is carried out in Ansys Maxwell, another important step in an automated

design optimization package is creating a well-defined parametric model which can

be implemented to reliably explore the entire design space.

In this case study design optimization, based on the overall shapes of the stator

and rotor structures, they can be parametrized as follows:

2.2.1.1 Parametrized Stator Structure

The stator geometry of the Prius propulsion motor is composed of common parallel

tooth and rounded slot bottom structures, which can be parametrized using the

geometric variables shown in Fig. 2.9.

Assuming the center of the machine axis to be the origin of the coordinate system,

the equations governing the principal nodes in this stator structure are expressed as

follows, see Fig. 2.9 for description of the parameters:


 p
xp1 = r2 − (wso /2)2
si
y = w /2
p1 so

xp2 = xp1 + dt2
y = y
p2 p1

xp3 = xp2 + dt3
y = y + d / tan(β) (2.2.1)
p3 p2 t3

xp4 = rsi + ds − rsb

wt
(rsi +ds ) tan( α2s )−
2 cos( α2s )
yp4 = rsb , where, rsb =

1+tan( α2s )

xp5 = rsi + ds
y = 0
p5
42

y
wt /2 4 x
1 23 rsb
αs 5
ds osb hy

dt3 rso

wtip/2
dt2 wso

rsi
β

Figure 2.9: Parametrized stator slot of the Prius motor.

y dq
wq/2
x 5 4
7 wrad
3
6
hpm

8
αpm
1 2
αp αpp dpm

wpm

rri

rro

Figure 2.10: Parametrized rotor layout of the Prius motor.


43

2.2.1.2 Parametrized Rotor Structure

The rotor structure of the Prius Gen 2 propulsion motor is also a common

configuration with single-layer V-type PM layouts. Assuming the center of the

machine axis as the origin of the coordinate system, the parametrized structure shown

in Fig. 2.10 can be defined by following nodes:


xq1 = xq2 − 0.75hpm sin (αpm /2)
y = 0
q1

xq2 = rro − dpm
y = 0
q2

xq3 = (rro − wrad ) cos(αpp /2)
y = (r − w ) sin(α /2)
q3 ro rad pp
   
xq4 = (rro − wrad ) cos αp /2 − arcsin
 wq
2(rro −wrad )
  
yq4 = (rro − wrad ) sin αp /2 − arcsin
 wq
2(rro −wrad )
 (2.2.2)
xq5 = xq4 − dq cos(αp /2)
y = y − d sin(α /2)
q5 q4 q p

xq6 = xq2 + 2wpm cos(αpm /2)
y = y + 2w sin(α /2)
q6 q2 pm pm

xq7 = xq6 − hpm sin(αpm /2)
y = y + h cos(α /2)
q7 q6 pm pm

xq8 = xq2 − hpm sin(αpm /2)
y = y + h cos(α /2)
q8 q2 pm pm
44

Table 2.1: Boundaries of the design variables defined over the parameterized cross-
section of the Prius motor.
Parameter(xi ) Description xi,min xi,max
ksi rsi /rso 0.6 0.7
hg Fig. 2.11 0.7 mm 2.5 mm
kwt wt /αs 0.35 0.75
kwtt wtip /(wso + wtip ) 0.3 0.8
kdpm dpm /dpm,max 0.25 0.50
kwpm wpm /wpm,max 0.80 0.93
kwq wq /wq,max 0.5 0.9
hpm Fig. 2.11 3.8 mm 9.0 mm
αpm Fig. 2.11 20 deg. 32 deg.
hy Fig. 2.11 13 mm 25 mm

2.2.1.3 Final Parametrized Model

The final parameterized cross-section of the Toyota Prius 48-slot 8-pole IPM motor

is constructed based on the parameterized stator and rotor cross-sections, see Fig.

2.11. Ten independent design variables are defined in its structure, 4 pertaining to

the stator geometry, 5 pertaining to the rotor geometry, in addition to the airgap

height. The geometric parameters are rationalized and confined according to Table

2.1 so as to avoid surface interference between the structures of various components

of the motor. As depicted in Fig. 2.12, the wide variations of the geometric design

parameters, from the lower to upper bounds is crucial for full exploration of the design

space.

With reference to the original design, the stator outer and the rotor inner

diameters are fixed to 260mm and 111mm, respectively. The active stack length

of the laminations and rotor PMs of each design candidate are scaled to produce the

nominal torque at the rated stator winding current density.


45

hy
rso
Stator
ds
Slot wt

wso hg
rsi

wq dpm
Rotor hpm
PM wpm

αpm
rri

Figure 2.11: The parameterized cross-section of the Prius motor comprising 10


independent design variables, see Table 2.1.

The calculation of the nominal torque is performed under maximum torque per

ampere (MTPA) operation for the given rated current density. The angle of the

current phasor for MTPA operation is determined by sampling the average torque

using a reduced number of static solutions, followed by interpolation of the average

torque profile versus the advance angle [129] of the current phasor as shown in Fig.

2.13. The MTPA operation is achieved at the angle where the torque reaches its peak.

2.2.2 Optimization Procedure and Results

The optimization was carried out over 60 generations each consisting of 80 members.

Two objectives were defined:


46

(a) Lower bounds of the design variables

(b) Upper bounds of the design variables

Figure 2.12: Range of variation of the design parameters defined over the Prius motor
topology.
47

40
Average T
30 Samples

Torque (Nm)
MTPA
20

10

0
90 112.5 135 157.5 180
Advanced angle (deg. el.)

Figure 2.13: Determination of current advanced angle for MTPA operation.

• Minimize the active material cost, AMC, as formulated in Eq. (2.2.3).

AM C = 24 · mpm + 3 · mcopper + msteel (2.2.3)

where the mass, m, is in kg and the steel cost is considered as the one-

unit reference in this normalized/per-unit formulation. It can be assumed

that the AMC is an approximate indication of the total cost of each motor

configuration, given that the manufacturing expenses of the same motor

topology with identical winding configuration is usually comparable for different

design solutions.

• Minimize the power losses consisting of the stator core and copper losses.

Furthermore, the two following constraints were considered for this optimization case

study to:

• Limit the torque ripple to less than 15%.

• Limit the degree of PM demagnetization to less than 70% of the

retentivity/remnant Br .
48

Table 2.2: Geometric design parameters of the counterpart optimal designs obtained
from the Prius motor design optimization at nominal load point.
Parameter G59M12 G20M62 Prius
ksi 0.66 0.64 0.60
hg (mm) 0.71 0.71 0.73
kwt 0.72 0.69 0.72
kwtt 0.34 0.47 0.36
kdpm 0.43 0.25 0.44
kwpm 0.93 0.93 0.88
kwq 0.66 0.73 0.86
hpm (mm) 4.22 5.55 6.50
αpm 31.67 25.24 30.15
hy (mm) 22.11 21.77 20.10
Stack length 106 95 84

The machine performance was optimized at a load point corresponding to a torque

of 400N m and a rotational speed of 1 500r/min. A current density of 23A/mm2 and a

slot fill factor of 0.47 has been reported for the original Prius design at this load point

[130, 131], which were maintained the same for all the design candidates throughout

this optimization process.

The optimization results are plotted in terms of objectives, and are color-coded

with respect to three performance metrics, namely torque ripple, minimum flux

density of the PMs, and power factor in Fig. 2.14(a) through (c), respectively. Some

observations could be made regarding the existing trade-offs between the objectives

and constraints in this specific case study. Aside from the conflicting relationship

between the active material cost and power losses, it can be seen that the machines

with higher costs tend to have a lesser PM demagnetization, and higher power factor.

This can be attributed to the larger amount of expensive magnet material used in

their constructions.
49

300 25

Active material cost (pu)

Torque ripple (%)


20
200
15

10
100
G59M12 5
G20M62 Prius
0 0
0 5 10 15 20 25
Stator power losses (kW)

(a) Color-coded for torque ripple

300 1
Active material cost (pu)

0.8

(T)
200
0.6

PM, min
0.4
100

B
G59M12 0.2
G20M62 Prius
0 0
0 5 10 15 20 25
Stator power losses (kW)

(b) Color-coded for PM minimum flux density

300 1
Active material cost (pu)

Power factor (−)

200 0.8

100 0.6
G59M12
G20M62 Prius
0 0.4
0 5 10 15 20 25
Stator power losses (kW)

(c) Color-coded for power factor

Figure 2.14: Results of the optimization of the Prius motor at the nominal load point
expressed in terms of the designated objectives.
50

Table 2.3: Comparison of nominal performance between counterpart designs obtained


from the Prius motor design optimization at nominal load point.
Index G59M12 G20M62 Prius
Harmonic Distortion Line-Line Terminal Voltage (%) 5.5 7.6 12.8
Harmonic Distortion Phase Terminal Voltage (%) 16.2 14.6 20.0
Torque Ripple (%) 10 13 18
Total Losses (W) 5763 6857 8148
Efficiency (%) 90.84 89.58 87.81
Power Factor (-) 0.82 0.76 0.69
3
Torque Per Rotor Volume (kNm/m ) 142.1 174.5 223.72
Rotor Peripheral Velocity (m/s) 13.85 13.41 12.60
Peak Airgap Flux Density (T) 1.95 1.91 1.96
Peak Stator Tooth Flux Density (T) 1.84 2.00 1.97
Peak Stator Yoke Flux Density (T) 1.55 1.61 1.68
Peak Rotor Yoke Flux Density (T) 1.00 1.08 1.78
Minimum PM Flux Density (T) 0.4 0.38 0.41

In Fig. 2.14, two optimized designs, “G59M12” and “G20M62” are marked, the

number after G indicates the generation and the number after M indicates the member

candidate design. Using the same simulation tools, the performance of the Prius

motor is also computed at this load operating point, and is compared with respect to

the design space in Fig. 2.14.

Accordingly, the Prius design can be recognized as an optimal design in terms

of active material cost. However, the efficiency is further improved in the other two

selected design candidates. In fact, design G20M62 exhibits both lower losses and

lower costs compared to the Prius design.

The three counterpart designs and their performance metrics are thoroughly

compared in Tables 2.2 through 2.5 , and Figs. 2.15 through 2.18. These metrics

are obtained using full-fledge time-stepping 2-D FEA in Motor-CAD for the same

operating point and at a temperature of 100◦ C for all components as was the case
51

Table 2.4: Distribution of power losses in the counterpart designs obtained from the
Prius motor design optimization at nominal load point.
Index G59M12 G20M62 Prius
Copper Loss (W) 5126 6204 7477
Magnet Loss (W) 5 6 6
Stator Back Iron Hysteresis Loss (W) 268 255 249
Stator Back Iron Eddy Loss (W) 24 22 22
Stator Tooth Hysteresis Loss (W) 196 231 252
Stator Tooth Eddy Loss (W) 38 39 42
Rotor Back Iron Hysteresis Loss (W) 18 29 26
Rotor Back Iron Eddy Loss (W) 0 0 0
Rotor Magnet Pole Hysteresis Loss (W) 85 66 70
Rotor Magnet Pole Eddy Loss (W) 5 4 3

Table 2.5: Mass of active components in the counterpart designs obtained from the
Prius motor design optimization at nominal load point.
Index G59M12 G20M62 Prius
Stator Lamination (kg) 21.2 19.3 17.6
Winding (kg) 4.1 5.0 6.0
Rotor Lamination (kg) 10.1 7.8 4.7
Magnet (kg) 1.28 1.13 1.25

throughout the optimization process.

According to Table 2.2, the split ratios, i.e. the ratio of stator inner to outer

diameter, of the optimized designs are higher than the original design, which, in

addition to the longer stack lengths of the optimized designs, has resulted in a lower

torque per rotor volume in the optimized designs, as listed in Table 2.3.

If the machine length has to be confined within a predefined value, the maximum

stack length can be introduced as an additional constraint in the optimization fitness

function. Yet, since the volume of the machine is oftentimes directly related to

the efficiency, the conflicting relationship between limitation of stack-length and

minimization of power losses will further narrow down the feasible design space.
52

The higher rotor peripheral velocity of the optimized designs in Table 2.3,

necessitates structural analysis of the rotor body to ensure that the bridges can

withstand the centrifugal stresses at the maximum operating speed. This mechanical

aspect of the design will be addressed in the last Chapter of this dissertation where

an analytical procedure is presented for calculation of these centrifugal stresses and

subsequent adjustment of the rotor bridges.

From Table 2.4 whereby the distributions of the losses are indicated for the

counterpart designs, it can be seen that the contribution of the rotor core losses

and PM losses to the overall core losses is marginal compared to that of the stator

core losses.

The masses of the main active components in the counterpart designs are listed

and compared in Table 2.5. It is interesting to note that the proportion of the mass

of the lamination steel to the masses of copper and PM is higher in the optimized

designs, whereas the copper masses and the PM masses are comparable to those

obtained for Prius.

The cross-sections of the three counterpart designs as well as the field plots under

the rated operating conditions are provided in Fig. 2.15, according to which a higher

degree of saturation is evident throughout the cross-section of the Prius design.

The waveforms of the line to line back EMF of the three counterpart designs are

shown in Fig. 2.16. Also the harmonic distribution of the back EMF waveforms

and the induced voltages are provided in Fig. 2.17(a) and (b), respectively. It is

interesting to note that the three designs have comparable distortions in their no-

load back EMF waveforms. However, the high order harmonics, namely 7th , 11th ,
53

(a) G59M12 (b) G20M62

(c) Prius

Figure 2.15: Cross-sections and rated field plots of the three counterpart designs
obtained from the Prius motor design optimization at nominal load point.
54

1 Line a−b
Line b−c

Line EMF (pu)


0.5 Line c−a
0

−0.5

−1
0 60 120 180 240 300 360
Rotor position (deg. el.)

(a)

1 Line a−b
Line b−c
Line EMF (pu)

0.5 Line c−a


0

−0.5

−1
0 60 120 180 240 300 360
Rotor position (deg. el.)

(b)

1 Line a−b
Line b−c
Line EMF (pu)

0.5 Line c−a


0

−0.5

−1
0 60 120 180 240 300 360
Rotor position (deg. el.)

(c)

Figure 2.16: Line to line back EMF of the counterpart designs obtained from the
Prius motor design optimization at nominal load point (a) G59M12, (b) G20M62,
and (c) Prius.
55

Harmonic magnitude (pu)


G59M12
0.8 G20M62
Prius
0.6

0.4

0.2

0
1 2 3 4 5 6 7 8 9 10 11 12 13 14
Harmonic order

(a) No-load (back EMF)

1
Harmonic magnitude (pu)

G59M12
0.8 G20M62
Prius
0.6

0.4

0.2

0
1 2 3 4 5 6 7 8 9 10 11 12 13 14
Harmonic order

(b) Full load

Figure 2.17: Harmonic content of the line to line terminal voltage of the counterpart
designs obtained from the Prius motor design optimization at nominal load point.

and 13th , are introduced to the induced voltage waveform of the Prius design to a

larger extent.

Similar to the back EMF waveforms, according to Fig. 2.18(a) the cogging torque

profiles of the two optimized designs are inferior to that of the Prius design. However,

as can be seen in Fig. 2.18(b), the optimized designs demonstrate a lower value of

peak torque ripple under rated operating conditions when compared to Prius design.
56

3
G59M12

Cogging torque (Nm)


2
G20M62
1 Prius
0
−1
−2
−3
0 2 4 6 8 10 12 14
Rotor position (deg. mech.)

(a) Cogging

60
G59M12
Torque ripple (Nm)

40
G20M62
20 Prius
0
−20
−40
−60
0 60 120 180 240 300 360
Rotor position (deg. el.)

(b) Full load

Figure 2.18: Torque ripple of the counterpart designs obtained from the Prius motor
design optimization at nominal load point.

As can be seen in the metrics shown in Tables 2.2 through 2.5 , and Figs. 2.15

through 2.18, the selected optimized designs outperform the original Prius design at

the nominal load point. However, when the power losses throughout the entire range

of operation are considered, as shown in Fig. 2.19, the Prius design features higher

efficiency in the extended speed region.

The procedure for calculation of the efficiency maps provided in Fig. 2.19,

which are obtained through 1 600 performance evaluations over sample load points
57

300
95
250

Efficiency (%)
Torque (Nm) 200
90
150
100 85
50
80
1000 2000 3000 4000 5000 6000
Speed (r/min)

(a) G59M12

300
95
250

Efficiency (%)
Torque (Nm)

200
90
150
100 85
50
80
1000 2000 3000 4000 5000 6000
Speed (r/min)

(b) G20M62

300
95
250
Efficiency (%)
Torque (Nm)

200
90
150
100 85
50
80
1000 2000 3000 4000 5000 6000
Speed (r/min)

(c) Prius

Figure 2.19: Efficiency maps of the counterpart designs obtained from the Prius motor
design optimization at nominal load point.
58

150 25
20

Torque (Nm)

Energy (kJ)
100 15
10
50
5
0
0 50 100 150 200 250 300
Speed (rad/s)

(a) UDDS

80
Torque (Nm)

100

Energy (kJ)
60

40
50
20

0 0
0 100 200 300
Speed (rad/s)

(b) HWFET

Figure 2.20: Load operating points of Toyota Prius propulsion PM motor obtained
from ADVISOR.

equidistantly distributed in the torque speed plane, will be described in Chapter 5.

Here, the purpose of this comparison is to recognize the fact that although the nominal

efficiencies of the optimized designs are higher than that of the Prius design, the high

efficiency contours of the optimized designs retract faster as the speed increases.

In many applications, including traction propulsion motors, maintaining high

performance throughout the entire operating range is imperative. Specifically in

the example case, the electric machine undergoes a wide range of speed and torque

variations both in constant torque and constant speed regions.


59

40 15

Torque (Nm)

Energy (kJ)
30
10
20
5
10

0
0 100 200 300 400
Speed (rad/s)

(a) UDDS

40
10
Torque (Nm)

Energy (kJ)
30

20 5
10

0 0
0 100 200 300
Speed (rad/s)

(b) HWFET

Figure 2.21: Load operating points of Honda Insight propulsion PM motor obtained
from ADVISOR.

To further illustrate this argument, the operating points of two traction

propulsion motors, Toyota Prius and Honda Insight are derived by modeling the two

vehicles in Advanced Vehicle Simulator (ADVISOR). Two US driving cycles, urban

dynamometer driving schedule (UDDS), and highway driving schedule (HWFET) are

considered to also underscore the variations of the operation conditions in reference

to the driving cycle.

It can be seen in Fig. 2.20 and Fig. 2.21 that not only do these motors operate at a

multitude of load points, but depending on the driving cycle, these operating points
60

can be far away from the rated/nominal load point. Even including the efficiency

improvement at the maximum torque of the corner speed in the fitness function, as

reported by some investigators [100], cannot accommodate the swarm of the load

operating points which can be located anywhere on the torque-speed plane.

In light of the above discussion, a high fidelity design optimization practice for

motors with wide ranges of operation requires accounting for the performance over

all load points. The realization of this goal is the main objective of the research

presented in this dissertation.

2.3 Summary

In this chapter, the fast and high fidelity design synthesis methodology using CE-FEA

and DE for large-scale design optimization of PM machines for rated operation was

reviewed. The principles of CE-FEA and DE were discussed and were implemented in

the design optimization of the Toyota Prius 2004 IPM motor. It was illustrated that

the design optimization at rated operation does not guarantee optimal performance

for the entire range of operation.


61

CHAPTER 3
EFFECTS OF AMPERE LOADING LEVEL ON
OPTIMAL DESIGN OF IPM MOTORS

In the previous chapter, the Toyota Prius Gen 2 interior permanent magnet (IPM)

motor was optimized for nominal (or peak) load point operation. It was shown

that although the losses of the selected optimized designs were reduced for the

targeted load point, the efficiency during the extended speed operation was degraded.

The question still remains as to whether and how the optimal design rules vary

for different loading conditions. Provided that the machine operating condition is

characterized by the magnitudes and rate of change of flux and ampere loading in the

core and in the windings, further clarifications concerning this matter can be made

by investigating the optimal design rules at various ampere loading levels. Indeed,

the design optimization of PM motors for a specific loading condition, which leads to

derivation/interpretation of optimal scaling rules for that particular loading condition,

has been widely investigated in the literature [87]. This chapter demonstrates how

these derivations vary with respect to the machine ampere loading and ferrous core

saturation level. For this purpose, a two-fold approach is adopted. First, a parallel

sensitivity analysis is carried out on three machines with three levels of ampere

loading. This is in order to gain an understanding as to how the correlation between

the design parameters and performance metrics vary under different ampere loading

conditions. Subsequently, a large-scale design optimization based on evolutionary


62

algorithms (EA), is pursued in order to identify how the optimal ranges of the design

variables are influenced by the Ampere loading and magnetic core saturation levels.

In this respect, ampere loading levels prevalent in three common cooling systems,

that is naturally cooled (NC), fan-cooled (FC), and liquid-cooled (LC) machines are

investigated. To fully account for the complex geometry, magnetic core nonlinearities,

and stator and rotor losses, a full-fledge time-stepping finite element(FE)-based

simulation platform is developed. Furthermore, to distinguish the intrinsic loss and

ripple characteristics of distributed and fractional slot concentrated winding machines,

in addition to the Toyota Prius 48-slot 8-pole configuration which was introduced in

the previous chapter, a 12-slot 10-pole configuration is also analyzed in this chapter.

3.1 Introduction

The design and modeling of PM motors has consistently been a subject of special

interest in the literature due to the distinctive features of such machines, including

but not limited to high efficiency, and high power density, which makes them suitable

for high performance applications [132, 133]. Recent design trends rely on large-

scale design optimization techniques based on electromagnetic (EM) finite element

(FE) analysis of design candidates for high fidelity calculation of their performance

metrics/characteristics [77]. In some studies, coupled EM/thermal models have been

utilized to accurately model the multi-physics nature of the design problem [89, 134–

136]. In many others, a fixed current density corresponding to the machine’s cooling

system, and accordingly an educated engineering guess of the operating temperatures


63

of various components are assumed [70, 137–141]. Although the coupled EM/thermal

models offer relatively reasonable predictions of temperature distributions inside such

machines, which can be used either to update the material properties for a subsequent

iteration of the EM FE analysis until a convergence is reached [89], or to determine

the maximum current density limit based on the winding temperatures [134], the

uncoupled approach is considerably faster. The uncoupled optimization approach

can also be as effective, provided that the design problem is well-defined and that the

thermal performances of the optimal designs are examined at the final stages of the

design process. The post-design optimization thermal modeling [142] is usually done

in order to identify the parameters of the cooling system such as the type of surfaces

or the coolant flow rate [143]. Since more efficient designs are achieved through the

design optimization, the cooling system is anticipated to remove lesser magnitudes

of losses than those associated with motors constructed on the basis of traditional

design methods which may yield suboptimal designs.

The design optimization practice is often accompanied by establishing the optimal

scaling rules for achieving a particular improvement in the machine performance.

Accordingly, design recommendations for mitigation of torque ripples, minimization

of power losses, and reducing dependency on rare-earth PMs, have been often reported

in the literature for naturally cooled machines [70, 85, 88, 138, 139]. Nevertheless,

increasing the stator winding current density or the machine’s ampere loading elevates

the flux level, and hence the saturation level throughout the machine’s magnetic

circuit. This not only affects the machine performance, but also alters the correlation
64

between them and the geometric design parameters associated with the machine cross-

section. In reference [144], the influence of ampere loading and magnetic saturation

on the cogging torque, back-emf and torque ripple of PM machines was investigated.

However, the impacts of the geometric design parameters on these performance

metrics was not considered. In reference [87], nonlinear scaling rules for low power

density brushless PM synchronous machines were developed without taking the effects

of ampere loading and magnetic core saturation into account.

This chapter contains further contribution through a parallel investigation of the

relations between the geometric design parameters and the performance metrics of

IPM motors for a wide range of ampere-loading levels determined by the machine’s

cooling system. This is accomplished through a systematic sensitivity analysis and

large scale design optimization. Accordingly, it will be demonstrated that one set of

design rules, such as those derived in [87], cannot be generalized to all three classes

of cooling systems.

In order to account for the distinctive performance characteristics of the integral

and fractional winding IPM machines, two mainstream case-study industrial IPM

machine configurations with distributed and fractional slot concentrated windings

(FSCW) are considered. In essence, while FSCW machines offer reduced copper

losses owing to shorter end windings [126, 145], they suffer from higher core losses

as a result of the space harmonics introduced by the more “discrete” nature of the

stator mmf waveform [146]. Furthermore, FSCW machines are less susceptible to

torque ripple due to their intrinsically lower cogging torque. This will lead to different

correlations of the geometric design parameters with these performance metrics as will
65

be discussed in this chapter.

3.2 Benchmark Designs

The parametric cross-sections of the studied machines and their full-fledged 2-D time-

stepping FE (TSFE) models are discussed in this section.

3.2.1 Parametrized Cross Sections

Two generic industrial IPM motor configurations with similar rotor layouts, though

with different slot-pole combinations are considered:

1. The first case-study is the 48-slot, 8-pole motor with a single-layer distributed

winding configuration, see Fig. 3.1(a). This is the same design configuration

as the Toyota Prius introduced and parametrized in the previous chapter. For

brevity, this design will be referred to as the 48S8P machine.

2. The second case study is a 12-slot, 10-pole motor with a so-called “double-layer”

concentrated winding configuration, see Fig. 3.1(b). The parametrization of the

stator design of this open-slot configuration is different than the 48-slot stator

design, and is elaborated in Appendix I. Hereafter, this design will be denoted

as the 12S10P machine.

The design parameters in Fig. 3.1 are rationalized according to Table 3.1 in order

to avoid geometric conflicts between the structures of various components of the two

motors. The parameterized FE model of the 48S8P motor is comprised of a total of

10 independent design variables, 5 residing in the rotor, 4 residing in the stator, in


66

hy
rso
Stator
ds
Slot wt

wso hg
rsi

wq dpm
Rotor hpm
PM wpm

αpm
rri
(a) The 48-slot 8-pole machine configuration

rso hy

Stator
Slot
ds wt

rsi hg
dpm
wq
Rotor hpm
PM wpm

αpm
rri
(b) The 12-slot 8-pole machine configuration

Figure 3.1: The parametrized cross section of the case study machines for sensitivity
analysis and optimization at different ampere loading levels.
67

Table 3.1: The independent design variables of the case study machines for sensitivity
analysis and optimization at different ampere loading levels.
48S8P 12S10P
Parameter(xi ) Description
xi,min xi,max xi,min xi,max
ksi rsi /rso 0.60 0.70 0.60 0.70
hg (mm) Fig. 3.1 0.70 2.50 0.70 2.50
kwt wt /αs 0.35 0.75 0.35 0.75
kwtt wtip /(wso + wtip ) 0.30 0.80 NA NA
kdpm dpm /dpm,max 0.25 0.50 0.25 0.50
kwpm wpm /wpm,max 0.80 0.93 0.80 0.93
kwq wq /wq,max 0.50 0.90 0.50 0.90
hpm (mm) Fig. 3.1 3.8 9.0 3.8 9.0
αpm (deg.) Fig. 3.1 20 32 19 26
hy (mm) Fig. 3.1 13 25 13 25

addition to the air-gap height. The 12S10P motor under investigation has one less

independent design variable due to its open slot configuration, thus eliminating the

parameter associated with the tooth tip.

The design variables are confined by upper and lower bounds, also listed in Table

3.1, either to prevent the unintended intersection of various boundary surfaces of

components or to address mechanical constraints, e.g. minimum air-gap height.

Nevertheless, wide bounds are designated to allow a full exploration of the design

space during the optimization process. For both machines, the stator outer and the

rotor inner (shaft) diameters are fixed to 260mm and 111mm, respectively.

3.2.2 Time-Stepping FE Models

The 2-D TSFE models of the two machines were developed in ANSYS Maxwell. The

core laminations were assumed to be non-oriented silicon steel of 0.36mm thickness.


68

The remenance and coercivity of the NdFeB PMs, and the resistivity of the copper

of the stator windings were evaluated corresponding to temperatures of 100 ◦ C, and

150 ◦ C, respectively. This is in order to account for the lowest magnet field and

highest winding resistance.

The TSFE analysis with sinusoidal current excitation is utilized to compute the

following performance metrics of interest:

1. Active material cost: The normalized active material cost (AMC) is calculated

using the same formulation provided in Eq. (2.2.3) in Chapter 2.

2. Power losses: The power losses consist of copper losses in the stator windings

with approximations of the end winding copper losses, core losses including

hysteresis and eddy current losses in the stator and rotor laminated cores,

and the eddy current losses in the rotor PMs for the 12S10P machine with

the assumption that these eddy currents are resistance limited. For accurate

calculation of the core losses, the frequency domain core loss model introduced

in [122] is utilized. The core losses have been obtained on an element by element

basis for an excitation current frequency corresponding to a rotor speed of

1800r/min.

3. Torque ripple: Following the calculation of the average torque per unit length,

Tave , using TSFE analysis, the stack-lengths of the designs are adjusted to

produce the desired average torque of 300 Nm for a stator current density

specifically chosen according to the cooling system of the machine. The

torque ripple, Tr , is subsequently determined from the torque profile over a


69

Table 3.2: Typical current density ranges found in electric machines with different
cooling systems.
Cooling Natural Fan Liquid
2
Current density (A/mm ) 1.5 − 5 5 − 10 10 − 30

full fundamental ac cycle using Eq. (2.1.4).

4. PM demagnetization: PM demagnetization effects are characterized by

the minimum flux density in the rotor PMs over a complete ac cycle.

Demagnetization is considered at the magnet piece level. Using NdFeB

PMs with reduced heavy earth dysprosium (Dy) content [147] can offer

local demagnetization protection, assuming that reasonable measures, such as

introducing air pockets around vulnerable areas [148], are adopted at the final

stages of the design.

The excitation current densities considered for the TSFE analysis accounts for

the type of the cooling system. Typical ranges of current densities are given in Table

3.2, [9]. Here, the fixed current densities of 4A/mm2 , 8A/mm2 , and 16A/mm2 are

assumed for naturally cooled (NC), fan-cooled (FC), and liquid-cooled (LC) machines,

respectively.

The phase angle of the current vector is chosen so as to ensure Maximum Torque

per Ampere (MTPA) operation. Since the design candidates can be saturated,

particularly in FC and LC classes, the torque angle at MTPA for each individual

design is numerically calculated by sampling and interpolating the generated torque

at multiple phase angles of the stator current phasor [70, 149], in a similar manner
70

to the procedure described in Chapter 2.

3.3 Parallel Sensitivity Analysis at Different

Ampere Loading Levels

The variation of the relationships between the geometric design parameters and the

performance metrics can be understood by carrying out a sensitivity analysis at

the three levels of stator winding current density. In the following subsections, the

sensitivity analysis methodology and the results are presented.

3.3.1 Methodology

For each cooling system, five second order response surfaces were defined

corresponding to the cost, copper losses, core losses, torque ripple, and degree of

PM demagnetization. The core losses and copper losses were treated separately since

the design variables have distinctive effects on these two loss components.

The regression coefficients pertaining to these response surfaces indicate how

the variation of the design parameters within their permissible ranges influences

the machine performance metrics. The class of Central Composite Designs (CCD)

[150] was used for design of experiments (DOE) in order to calculate the regression

coefficients associated with each second order response surfaces, y, given in Eq.

(3.3.1),
n
X n
X n X
X n
y = βo + βi ci + βii c2i + βij ci cj (3.3.1)
i=1 i=1 i=1 j=i+1

where βo , βi , βii , and βij are the regression coefficients for the n design variables, xi ,
71

i = 1, . . . , n expressed in the coded form, ci , according to Eq. (3.3.2).

xi − (xi,max + xi,min )/2


ci = , i = 1, 2, ..., n (3.3.2)
(xi,max − xi,min )/2

In Figs. 3.2 and 3.3, the regression coefficients of the sensitivity analysis are shown

for the five examined performance metrics at the three different levels of current

densities accounting for the three classes of cooling systems. Since the purpose of the

sensitivity analysis is to provide a measure of importance of each design parameter

with respect to other design variables, the regression coefficients are normalized to

the variable with the maximum influence in each group of metrics. Furthermore,

the DOE is conducted so as to ensure no main effect or two-factor interaction is

aliased with any other main effect or two-factor interaction of the design variables,

i.e. so-called resolution V designs [150]. Therefore, the regression coefficients can be

examined independently for each group of metrics.

Each coefficient is associated with a design variable, i.e. input, to indicate how

its variation within the permissible range denoted in Table 3.1 would influence the

machine’s performance, i.e. output. A positive coefficient, (+), indicates that an

increase in input will raise the output whereas a negative coefficient, (−), constitutes

the opposite trend between the input and output. Accordingly, a close examination

of Figs. 3.2 and 3.3 leads to the observation that some trends vary drastically, or even

change direction, with respect to the ampere loading for the two 48S8P and 12S10P

machines.

3.3.2 Discussion of the Results of the Sensitivity Analysis

The sensitivity analysis for each performance metric is summarized in this section.
72

NC FC LC
1

Cost 0

−1
(a)
1
Copper loss

−1
(b)
1
Core loss

−1
(c)
1
Ripple

−1
(d)
1
BPM,min

−1
tt
ksi

pm

αpm
hg

pm

pm

hy
q
kw
w

kw
k

kd

kw

(e)

Figure 3.2: Normalized regression coefficients from the sensitivity study indicating
the effect of the design variables on, (a) active material cost, (b) copper loss, (c) core
loss, (d) minimum PM flux density, and (e) torque ripple for the 48S8P machine.
73

NC FC LC
1

Cost 0

−1
(a)
1
Copper loss

−1
(b)
1
Core loss

−1
(c)
1
Ripple

−1
(d)
1
BPM,min

−1
ksi

pm

pm
hg

pm

pm

hy
q
w

kw
k

kw

α
h
k

(e)

Figure 3.3: Normalized regression coefficients from the sensitivity study indicating
the effect of the design variables on, (a) active material cost, (b) copper loss, (c) core
loss, (d) minimum PM flux density, and (e) torque ripple for the 12S10P machine.
74

3.3.2.1 Active Material Cost

According to Figs. 3.2(a) and 3.3(a), for both machine configurations:

• The strongest correlation with AMC in the NC and FC classes is attributed to

the yoke height, which is opposite to the slot depth for a given stator bore.

• The yoke height correlation with AMC is superseded by that of the air-gap

height in the LC class.

• The positive correlation of the tooth stem width with AMC monotonically

decreases as the ampere loading level increases.

3.3.2.2 Copper Losses

According to Figs. 3.2(b) and 3.3 (b):

• The considerable variations of the influences of the design parameters on the

copper losses as the ampere loading level increases should be noted for both

machine configurations.

• For the 48S8P machine, the air-gap height constitutes the strongest correlation

with copper losses in the NC class. In the FC and LC classes, the split ratio,

ksi , the tooth-stem width, and the yoke height become more influential.

• Similar trends exist for the 12S10P machine, except for the diminished influence

of the yoke height.


75

3.3.2.3 Core Losses

According to Figs. 3.2(c) and 3.3 (c):

• The influences of the design parameters on the core losses are sensitive to the

ampere loading to such an extent that some of these relationships, i.e. those

associated with the stator inner diameter and yoke height, are reversed as the

ampere-loading increases.

• For both machine configurations, in the NC class, the design parameters

associated with the stator constitute the strongest correlation with the core

losses.

• As the ampere loading increases, the design parameters associated with the

rotor become more influential.

• The strong influence of the rotor q-axis bridge on the core losses, especially in

the 12S10P motor, should be noted.

3.3.2.4 Torque Ripple

According to Figs. 3.2(d) and 3.3(d):

• Except for the air-gap height which maintains a strong negative correlation

with torque ripple under any ampere loading conditions, the two machine

configurations have distinctive torque ripple characteristics.

• For the 48S8P machine, the width of the q-axis bridge becomes more influential

than the air-gap height as the level of ampere-loading increases. Meanwhile, the
76

negative correlation of the widths of the tooth stem and tooth tips with torque

ripple diminishes as the saturation level of the magnetic core is elevated.

• As opposed to the 48S8P machine, the 12S10P machine is intrinsically less

susceptible to torque ripple under any loading conditions.

• It is interesting to note the reversal of the correlation between the torque ripple

and the tooth stem width under heavy magnetic core saturation in the 12S10P

motor.

3.3.2.5 PM Demagnetization

According to Figs. 3.2(e) and 3.3(e):

• For both machine configurations and under all loading conditions, the PM

demagnetization is strongly correlated to the PM height and tooth-stem width.

• The correlation factors of the split ratio, ksi , and the yoke height monotonically

increase as the level of ampere loading increases.

• Although the air-gap height constitutes a negative correlation with Bpm in NC

machines, this relationship reverses as the ampere-loading level increases.

3.4 Parallel Statistical Analysis of the Optimized

Designs at Different Ampere Loading Levels

The sensitivity analysis merely reveals the independent effects of the design variables

on the machine performance. To understand how the interactions between these


77

variables influence the final optimal design which is subject to a unique set of

objectives and constraints, a close examination of the design optimization process

and results is in order. Nevertheless, the sensitivity study suggests that the optimal

designs for each class should have distinctive features which will be examined in this

section.

A large-scale design optimization algorithm has been developed and utilized to

optimize the two motor configurations for each level of the current densities. The

design optimization is followed by a statistical analysis on the optimized designs to

find the range of change, i.e. the statistical distribution of the design variables in the

optimized designs residing in the design space for each class of the cooling systems.

3.4.1 Large-Scale Design Optimization

Six runs of a large-scale design optimization, each composed of 6600 candidate designs,

were carried out to optimize the machine model at the three aforementioned levels of

current densities.

The optimization algorithm relies on the TSFE model developed in Section 3.2.

A Combined Multi-objective Optimization with Differential Evolution (CMODE)

[110], was implemented as the optimization search algorithm. This CMODE-type

optimization is fully described in Chapter 4.

The fitness function of the optimization problem has been defined based on the

previously discussed performance metrics. For generic industrial use, it practically

consists of the two following objectives:

• Minimization of the AMC given in Eq. (2.2.3)


78

1000
NC

AMC (per unit steel)


800 FC
LC
600

400

200

0
0 2000 4000 6000 8000 10000
Power losses (W)

(a) 48S8P design

1000
NC
AMC (per unit steel)

800 FC
LC
600

400

200

0
0 2000 4000 6000 8000 10000
Power losses (W)

(b) 12S10P design

Figure 3.4: Optimization results for the three case-study cooling systems.

• Minimization of the power losses

Due to the opposite correlations of the design parameters, except the air-gap

height, hg , with AMC and power losses in Figs. 3.2(a), and 3.3(a), the two objectives

are conflicting.

Furthermore, the two following constraints are introduced for reliable operation:

• Less than 15% torque ripple

• Less than 70% PM demagnetization


79

Each run of the optimization was started with an initial generation of 200 members

and was carried out over 800 generations of 8 members each [110]. Figures. 3.4(a)-

(b) show the optimization results in terms of the conflicting objectives for the 48S8P

and 12S10P machines, respectively. The design candidates indicated in Fig. 3.4 are

compared in terms of AMC and power losses at the same rated load operating point.

It is evident that increasing the stator winding current density, and thus the

ampere loading of the machine, complicates the design of the cooling system by

increasing the total dissipated losses. Nevertheless, in many applications, higher

torque density can translate into reduced system cost, size, and weight, thus

promoting such designs with sophisticated cooling systems [142, 148].

Furthermore, the extremes of the power losses and AMC of the three classes of

cooling systems are non-overlapping. That is, the same efficiency characteristics of NC

machines cannot be obtained by LC machines. Similarly, the AMC in LC machines

can be reduced to values not achievable by NC machines.

It is interesting to notice that due to higher core losses and larger masses of PM in

their construction, the Pareto-optimal designs of the 12S10P configuration are located

further apart from the origin as compared to the 48S8P configuration, see Tables 3.4

and 3.5.

3.4.2 Scaling Rules of the Optimum Candidate Designs

To differentiate the distinctive optimal design values between the three cooling classes,

the statistical distributions of the design variables in the Pareto-optimal designs

should be investigated. For this purpose, from each run of optimization, 500 superior
80

xi,max
LC
FC
NC

x
i,mid

x
i,min

tt
ksi

hpm

αpm
g

pm

pm

hy
q
kw
w

kw
h

kd

kw
(a) 48S8P design

xi,max
LC
FC
NC

xi,mid

xi,min
ksi

hpm

αpm
g

pm

pm

hy
q
kw

kw
h

kd

kw

(b) 12S10P design

Figure 3.5: The distribution of the design variables in the optimized designs for the
three case study cooling systems.

designs were selected based on a three-step process. First, the designs which do

not violate the constraints on the torque ripple and on the PM demagnetization are

identified and separated in set P .

Subsequently, a strength value s(di ) is designated to each of the designs, di in P ,


81

according to Eq. (3.4.1).

s(di ) = #{dj | dj ∈ P and cost(di ) < cost(dj )


(3.4.1)
and loss(di ) < loss(dj ) }, i = 1, 2, ..., k

where ”#” is the cardinality of the set [151], and k is the number of the designs in P .

The designs in P are thereafter ranked in a descending order based on their strength

value, sdi , to determine the top 500 superior designs.

The distribution of the design parameters in the selected designs with respect to

their bounds can be described by “box plots” shown in Fig. 3.5. The rectangular

boxes in Fig. 3.5 represent the first, the second and the third quartiles of the

distribution of design values with respect to the upper and lower bounds designated

for each design parameter in Table 3.1. The distances between the different parts

of these boxes indicate the degree of dispersion and skewness in the value of the

optimal designs. The circles in the middle, and the whiskers represent the average,

the maximum and the minimum of the optimal design value for each parameter.

The representative cross-sections which were generated based on the means of the

design variables in Table 3.3 are provided in Fig. 3.6 which allows one to visualize

the distinctive design features of each machine. These features will be subsequently

explained with reference to the statistical and sensitivity analyses.

According to Fig. 3.5, the optimal ranges of the design variables are broader in

the 48S8P machine. This stems from the imposed criterion on the torque ripple of

the selected designs, which leads to the dispersion of the chosen design candidates in

the Pareto-front vicinity.

Three different trends can be recognized in the variation of the design parameters
82

(a) 48S8P design

(b) 12S10P design

Figure 3.6: Optimized cross sections derived based on the mean of the design variables
in the 500 Pareto-optimal designs listed in Table 3.3, for the three current density
levels.

Table 3.3: Average design parameters of the optimized designs at different ampere
loading levels.
48S8P 12S10P
Parameter(xi )
NC FC LC NC FC LC
ksi 0.62 0.64 0.66 0.63 0.67 0.69
hg (mm) 0.83 0.75 0.72 0.74 0.71 0.71
kwt 0.52 0.61 0.66 0.41 0.57 0.67
kwtt 0.63 0.67 0.57 NA NA NA
kdpm 0.41 0.42 0.43 0.48 0.49 0.48
kwpm 0.91 0.92 0.92 0.47 0.46 0.46
kwq 0.53 0.56 0.57 0.53 0.53 0.52
hpm (mm) 6.71 5.06 4.22 8.38 7.40 5.55
αpm (deg.) 31.73 31.21 31.21 25.50 25.74 25.73
hy (mm) 16.52 17.72 20.12 14.63 14.51 17.40

Table 3.4: Mean of the ratio of copper losses to core losses in the Pareto-optimal
designs at different ampere loading levels.
Machine NC FC LC
48S8P 2.16 5.57 15.24
12S10P 0.74 1.52 3.47
83

Table 3.5: Mean of the masses of the optimized designs normalized with respect to
the values obtained for NC class.
48S8P 12S10P
Mass (m)
NC FC LC NC FC LC
mP M % 100 47.47 26.22 100 56.53 32.60
mcopper % 100 52.14 30.22 100 45.10 23.31
msteel % 100 59.88 38.34 100 58.14 42.32

which are addressed in the following discussion.

3.4.2.1 Parameters With Increasing Trends in Both Motor Configura-

tions

As can be seen in Fig. 3.5, the split ratio, ksi , the ratio kwt , and the yoke height,

hy , relatively grow with the increase of the ampere loading. This is in line with

the sensitivity analysis and the fact that decreasing copper losses, as the major loss

component in the case study PM machines according to Table 3.4, takes precedence

over decreasing core losses. The kdpm ratio varies widely in the 48S8P machine due to

the insignificant correlation of this design variable with the optimization objectives

according to the sensitivity analysis in Fig. 3.2. The stronger negative correlation of

kdpm ratio with AMC and copper and core losses in 12S10P machine, has resulted in

its maximization in the selected 12S10P designs, as shown in Fig. 3.5(b). Similarly,

the kwpm ratio and αpm are concentrated towards their higher bounds due to their

negative correlations with the two objectives as shown in Figs. 3.2 and 3.3.
84

3.4.2.2 Parameters With Decreasing Trends in Both Motor Configura-

tions

The PM height, hpm , in Fig. 3.5 monotonically decreases as the ampere loading

increases in both 48S8P and 12S10P machines. This can be explained by considering

the results of the sensitivity analysis in Figs. 3.2 and 3.3, and the change of the masses

of various components in the selected designs for the three cooling systems, as listed

in Table 3.5. Accordingly, the PM mass has been reduced in the selected designs

proportionately to the masses of other components by reducing the only parameter

associated with the PM mass which constitutes a positive correlation with AMC, i.e.

hpm . Meanwhile, the excessive PM demagnetization as a result of reducing PM height

has been prevented by introducing a constraint on the minimum flux density of the

PMs. The air-gap height, hg , in the optimum designs is minimized due to its positive

correlation with AMC and major component of the losses, i.e. copper losses.

3.4.2.3 Parameters With Conflicting Trends in the Two Motor

Configurations

For the 48S8P machine, while the kwq ratio monotonically increases in Fig. 3.5(a)

as the torque ripple becomes less of an issue, it decreases constantly in the 12S10P

machine, Fig. 3.5(b), in order to reduce the core losses. On the one hand, according

to the sensitivity analysis, the core loss has a stronger correlation with this design

variable, and on the other hand, the core losses constitute a considerable portion of

the overall losses in the 12S10P machine, see Table 3.4.

The tooth-tip width of the 48S8P machine widely varies at higher ampere loading
85

due to saturation. The tooth-tips are larger at low ampere loading to alleviate

the torque ripple, see Fig. 3.2(d). Their variation is ineffective when the machine

saturates. Meanwhile, the mean of the tooth-tip width in Fig. 3.5(a) is reduced to

mitigate the core losses, although this reduction in width is not significant since the

overall ratio of core to copper losses is insignificant, see Table 3.4.

3.5 Summary

A parallel sensitivity analysis was carried out on two case-study IPM machines with

concentrated and distributed stator winding configurations and with different cooling

systems. It was demonstrated that the correlation between the main design variables

and various performance metrics, particularly core losses, copper losses, and torque

ripple can be significantly affected by the machine’s ampere loading and magnetic

core saturation. In some cases, these trends can even be reversed.

The distribution of the optimal design values were investigated for each case-

study in a practical optimization problem where the interaction of the performance

metrics and design variables occur. Noticeable difference in the optimal design values

were observed and the trends were classified for the naturally cooled, fan-cooled, and

liquid-cooled machines.

The results of this chapter accentuate the challenges in the design of electric motors

with intermittent operating cycles, such as those in traction applications, where an

optimal design is to maintain high performance under various loading conditions, as

will be dealt with in following chapters.


86

CHAPTER 4
MULTIOBJECTIVE CMODE-TYPE
OPTIMIZATION OF ELECTRIC MACHINES

In the two previous chapters, it was demonstrated that the optimal design rules of

PM machines vary with respect to the ampere loading conditions. As a result, for

design synthesis of PM machines with a wide range of operation, these ampere loading

conditions should be accounted for in the optimization fitness function. As will be

discussed in the following chapters, one method to do so is by evaluating the machine

performance at multiple load operating points which represent various loading

conditions of the machine. This, however, increases the computational complexity

of the overall design optimization process, even if computationally efficient modeling

approaches such as CE-FEA are used. To further improve the computational

efficiency, a new optimization algorithm is developed and presented in this chapter,

which can outperform the existing DE optimizer when implemented on multi-core

desktop computers. As shown in Chapters 2 and 3, large-scale design optimization

of electric machines is oftentimes practiced to achieve a set of objectives under a

set of constraints. Accordingly, the design optimization of electric machines can

be regarded as a constrained optimization problem (COP). Evolutionary algorithms

(EAs) used in the design optimization of electric machines including DE, which has

received considerable attention during recent years, are unconstrained optimization

methods that need additional mechanisms to handle COPs. In this chapter, a new
87

optimization algorithm that features combined multi-objective optimization with

differential evolution (CMODE) has been developed and implemented in the design

optimization of electric machines. A thorough comparison is conducted between

the two counterpart optimization algorithms, CMODE and DE, to demonstrate

CMODE’s superiority in terms of convergence rate, diversity and high definition of the

resulting Pareto fronts, and its more effective constraint handling. More importantly,

CMODE requires a lesser number of simultaneous processing units which makes its

implementation best suited for state-of-the-art desktop computers reducing the need

for high performance computing (HPC) systems and associated software licenses.

4.1 Introduction

Large-scale design optimization techniques have become a well-established practice

for designing high performance electric machines [73, 84, 85, 96, 119, 129, 152, 153].

In these techniques, the parametrized cross-section of a subject machine is refined to

improve certain performance metrics with respect to the application requirements.

Cost, power loss, torque density, torque ripple, power factor, and degrees of

demagnetization of rotor PMs in PM machines, are common performance metrics

which constitute the set of objectives and constraints in any optimization problem.

In principle, a large-scale model-based design optimization process consists of two

independent segments: the machine model for computation of performance metrics,

and the optimizer for finding the globally optimal design solutions. Regarding

the machine model, both analytical [73, 129, 152, 153] and Finite Element (FE)
88

[84, 85, 96, 119] methods are commonly used in a large-scale design optimization

process, with the latter receiving more attention during recent years owing to the

ever-increasing processing power of modern computers. Concerning the optimizer,

either a deterministic or stochastic search algorithm can be used.

Efficient utilization of computational resources is imperative when the

performance evaluation of the design candidates is computationally intensive [78], as

in the case of the FE models. Two areas of opportunity exist under each segment of

the optimization process that can serve this purpose. On the performance evaluation

side, CE-FEA has been recently introduced [79–81, 84] for fast and high fidelity

simulation of PM machines. On the optimizer side, the DE algorithm [154] has

received extensive attention as a reliable and fast stochastic search algorithm [77, 155].

The DE is thought to have a better performance in comparison with other stochastic

optimizers in electric machinery design problems [77, 120]. It has been coupled to

the CE-FEA for optimization of several types of PM motors with various sets of

objectives and constraints [84–87, 96, 119].

Although DE has proved effective in the design optimization of electric machines

[94, 156], similar to other EAs, it has not been developed for handling COPs [151, 157],

which is the case in design of electric machines [94, 156, 158]. Popular constraint

handling mechanisms include penalty function methods, methods based on preference

of feasible solutions, and multi-objective optimization techniques [151, 158]. In the

latter, COPs are converted to unconstrained multi-objective optimization problems

where minimization of the so-called degree of constraint violation is designated as an

additional objective.
89

In this Chapter, a recently developed CMODE algorithm [110, 151] is adapted for

the design optimization of electric machines with application to three IPM motors

with distributed and concentrated stator windings. The same design problem is

performed using the standard DE to compare the outcomes with those obtained

from the CMODE approach. It is demonstrated here that CMODE is superior

to DE in terms of convergence rate and constraint handling in all three example

motor configurations. Furthermore, CMODE requires a lesser number of simultaneous

function evaluations which makes it an attractive solution for implementation of the

design optimization on a state-of-the-art desktop computer with a limited number of

processors, thus reducing the need for high performance computing (HPC) facilities

and associated software licenses. First, the essence of the two optimization algorithms

with a focus on their similarities and differences are discussed. Subsequently, following

the description of the benchmark studies, the optimization results and the comparison

between the two algorithms are provided.

4.2 CMODE Versus DE in the Design Optimiza-

tion of Electric Machines

The flowcharts of steps of the two counterpart search algorithms, DE and CMODE,

applied to the design optimization of electric machines are shown in Fig. 4.1 and Fig.

4.2, respectively. In both cases, an initial design is obtained analytically in reference to

the application requirements and specifications [9]. This initial design is subsequently

parametrized and the geometric design variables and constants are specified in the
90

Start large-scale DE
optimization algorithm

● Enter the boundaries of the


design variables.
● Initialize mutation and
crossover operations

● Randomly generate an initial population (P) of size NP


from the decision space as the first generation (g).
● Evaluate the performance indices of each individual in
P using CE-FEA.
● Number of function evaluations (FEs ) = NP, g = 1.

● For each individual in Q, an offspring is generated by using the


mutation and crossover operations of DE resulting in trial set T.
● Evaluate the performance indices of each individual in C using
CE-FEA.
● Identify the surviving individuals based on Lampinen’s
selection criteria.
● Update P, g = g+1, FEs = FEs + NP.

No
FEs > FEsMAX

Yes
Stop

Figure 4.1: Flowchart of the steps of the DE optimization algorithm.

initialization stage. Preparation of a well-defined parameterized model, which on the

one hand is flexible for the exploration of the entire design space, and on the other

hand is restrained to avoid geometric conflicts between various components of the

machine cross-section, is a non-trivial demanding task.

The optimization process involves performance assessment of the resulting design

candidates. For this purpose, the CE-FEA approach [80, 84] described in Chapter

2, which accommodates the complex geometry of the machine structure, and

incorporates the actual non-linear nature of the magnetic core, is utilized.


91

Start large-scale CMODE


optimization algorithm

● Enter the boundaries of the


design variables.
● Initialize NP, η, k, FEsMAX

● Randomly generate an initial population (P) of size NP


from the decision space as the first generation (g).
● Evaluate the performance indices of each individual in
P using CE-FEA.
● Number of function evaluations (FEs ) = NP, g = 1.

● Choose η individuals (set Q) from P randomly.


● For each individual in Q, an offspring is generated by using the
mutation and crossover operations of DE resulting in set C.
● Evaluate the performance indices of each individual in C using
CE-FEA.
● Identify the non-dominated/superior individuals (set R) from C.
● Each individual in R randomly replaces an individual in Q that
is dominated by/superior to it to update Q.
● If R only contains infeasible solutions then the infeasible
solution with lowest degree of constraint is stored into archive A.
● Update Q in P, g = g+1, FEs = FEs + η.

No
Reminder (g, k) = 0

Yes
● Randomly choose one of the following infeasible
solution replacement methods: deterministic or random.
● In deterministic replacement, the individuals in P are
ranked based on their quality and feasibility. The
individuals with the lowest rank are then replaced with the
individuals in A.
● In random replacement, the individuals in P are
randomly selected/replaced with the individuals in A. The
replacement does not apply to the best individuals of P.

No
FEs > FEsMAX

Yes
Stop

Figure 4.2: Flowchart of the steps of the CMODE optimization algorithm.


92

Apart from using the same parameterized model and the same technique for

performance evaluation of the design candidates, here the CE-FEA method, the

optimization procedure differs for the CMODE and the DE algorithms in the following

manner:

• Unlike DE in which, as a standard EA, all the population members are

used to generate the offspring population, in CMODE, only a portion of the

total individuals, denoted by η, are chosen for this purpose. This renders

CMODE a steady-state EA where the first randomly generated population has

a large number of members, a fraction of which, set Q, are being constantly

updated throughout the optimization process. Consequently, CMODE performs

a fraction of simultaneous function evaluations (FEs) contrary to what takes

place in DE. Here, performing FEs means applying the CE-FEA approach in

solving the electromagnetic field in the design candidate machines. Typical

numbers recommended for DE are 60 generations each consisting of 80 members

[85, 96], in contrast with an initial population of 180 members followed by

400 generations each consisting of 8 members recommended for CMODE

[151]. Lesser number of simultaneous FEs, here 8 versus 80, makes CMODE’s

implementation best suited for state-of-the-art desktop computers.

• The selection procedure for determination of the surviving candidates in

CMODE is based on the identification of superior individuals in the trial

population, set C, and having them replace the dominated individuals in the

parent population, set Q. Therefore, in comparison to DE, there is an additional


93

round of competition in CMODE. The first round is between all the individuals

in the trial population, and the second round is between the winners of the

first round, set R, and the individuals in the parent population. Nevertheless,

CMODE still benefits from the mutation and crossover operations of DE [151],

which produces the trial and ultimately the offspring populations of consecutive

generations.

• In CMODE, an additional variable defined as the degree of constraint violation

is introduced into the objective function to be minimized with other objectives.

Let ~x be the design vector, and f (~x) be the initial objective of the optimization

problem subject to a set of q inequality constraints, gj (~x), and (m − q) equality

constraints, hj (~x). The final objective of the optimization process would be the

minimization of f (x̃) = (f (~x), G(~x)), where G(~x) is the degree of constraint

violation given in Eq. (4.2.1) below [151].

m
X
G(~x) = Gj (~x) (4.2.1)
j=1

max{0, gj (~x)}, 1 ≤ j ≤ q
where Gj (~x) = {
max{0, | hj (~x) | −δ}, q + 1 ≤ j ≤ m

• Finally, according to Fig. 4.2, CMODE features an infeasible solution

replacement mechanism in which after a certain number of generations, denoted

by k, an archive consisting of individuals that violate the constraints, A, replaces

the individuals in the main population, P, either through a deterministic

or random procedure. This mechanism adds to the diversity of the overall

optimization problem to facilitate convergence to the globally optimal solutions.


94

4.3 Benchmark Study- Application to IPM Mo-

tors

4.3.1 Parametrized FE Model

To compare the merits of the two counterpart stochastic search algorithms, three IPM

motor configurations with distinctive rotor and stator features, and under different

loading conditions have been investigated:

1. A fan-cooled 48-slot, 8-pole motor with single-layer v-shaped magnets and

single-layer distributed winding configuration. This design will be referred to

as the 48S8P-a design for brevity.

2. A liquid-cooled 48-slot, 8-pole motor with double-layer v-shaped magnets and

single-layer distributed winding configuration, which will be referred to as the

48S8P-b design.

3. A naturally cooled 12-slot, 10-pole motor with single-layer v-shaped magnets

and a so-called “double-layer”, i.e. all teeth wound” concentrated winding

configuration, called hereafter the 12S10P design.

The current density of the stator winding can be adjusted to account for the

ampere-loading of the machine. Typical current density ranges are provided in Table

3.2 [9]. Here, 22 A/mm2 , 8 A/mm2 , and 4 A/mm2 are assumed for liquid-, fan-,

and naturally cooled machines, respectively. The variety introduced to the selected

machine configurations and their electrical loadings provides the basis for a rigorous
95

comparison between the two search algorithms.

To construct the FE model of the example machines, the parametrized stator

and rotor structures, respectively shown in Figs. 4.3 and 4.4 are utilized. The

independent design variables defined based on theses parametrized models are listed

in Table 4.1. Some of these design variables are rationalized according to Table 4.1 so

as to avoid geometric conflicts between the structures of various components of the

motor. The parameterized FE model of the 48S8P-a motor is comprised of a total

of 10 independent design variables, 5 residing in the rotor, 4 residing in the stator,

in addition to the air-gap height. The 48S8P-b motor has 2 additional independent

variables introduced to the rotor geometry to accommodate the double-layer PMs.

Meanwhile, the open-slot 12S10P motor has one less independent design variable

because of its open slot structure. The design variables are confined by upper and

lower bounds, also listed in Table 4.1 and depicted in Figs. 4.3 and 4.4 for some of the

variables in a typical design, either to prevent the unintended intersection of various

boundary surfaces of machine components, or to address mechanical constraints, e.g.

minimum air-gap height or the yield stress for the rotor bridges [159]. For all the three

machines, the stator outer diameter is fixed to 260mm. The shaft diameter is equal to

111mm, and 74mm for the single-layer and double-layer rotor magnet configurations,

respectively. The parameterized geometry together with the introduced bounds allow

the model to be flexible in exploring the entire design space to find the globally

optimized design candidates.

The CE-FEA method is utilized for fast and high fidelity calculation of the

machine performance metrics [80, 84]. This method was explained in Chapter
96

rso
rsi
rro

αwt
y
hy
αs ds
o x

wtip/2
wso wt

hg

(a) 48-slot stator

rso
rsi
rro

αwt
y
αs
o x ds hy

wt
hg

(b) 12-slot stator

Figure 4.3: The parameterized stator structures used for constructing the example
IPM motors for comparison between DE and CMODE.
97

hpm

wq,max
wq

rri

y
d'pm dpm
αpm
o x

wpm,max
wpm
rro

(a) Single-layer rotor PM layout

wq,max
wq

rri
hpm1

y hpm2
d'pm dpm
αpm1 αpm2
o x
wpm2

wpm1
wpm1,max hpm2,max

rro

(b) Double-layer rotor PM layout

Figure 4.4: The parameterized rotor structures used for constructing the example
IPM motors for comparison between DE and CMODE.
98

Table 4.1: Independent design variables of the parametric stator and rotor structures
shown in Figs. 4.3 and 4.4.
48S8P-a 48S8P-b 12S10P
Parameter(xi ) Description
xi,min xi,max xi,min xi,max xi,min xi,max
ksi rsi /rso 0.6 0.7 0.6 0.7 0.6 0.7
hg (mm) Fig. 4.3 0.7 2.5 0.7 2.5 0.7 2.5
kwt αwt /αs 0.35 0.75 0.35 0.75 0.35 0.75
kwtt wtip /(wso + wtip ) 0.3 0.8 0.3 0.8 NA NA
0
kdpm dpm /(dpm + dpm ) 0.25 0.50 0.25 0.50 0.15 0.65
kwpm wpm /wpm,max 0.80 0.93 0.80 0.93 0.76 0.94
kwq wq /wq,max 0.5 0.9 0.65 0.90 0.3 0.8
hpm (mm) Fig. 4.4 3.8 9.0 3.8 9.0 2.5 8.0
αpm (deg.) Fig. 4.4 20 32 20 32 19 26
hy (mm) Fig. 4.4 13 25 13 25 13 25
kαpm αpm1 /αpm2 NA NA 0.3 0.8 NA NA
khpm hpm2 /hpm2,max NA NA 0.4 0.8 NA NA

2. Using the CE-FEA allows reconstruction of the entire field waveforms through

multiple snapshots of magnetostatic FE solutions over a time span corresponding

to 60 electrical degrees. The CE-FEA can be up to two orders of magnitude faster

compared to the full-fledged time-stepping transient FE solutions [80, 84]. It has

been demonstrated to be effective in large-scale design optimizations of PM machines

with various rotor layouts and stator winding configurations, including experimental

verifications [84, 85, 87, 96, 119].

4.3.2 Optimization Fitness Functions

Since the purpose of the optimization is a comparative study between the search

algorithms, the fitness function of the optimization problem can be chosen arbitrarily.

In a practical case, the performance metrics of interest can be the machine’s active
99

material cost, power losses, torque ripple, and the degree of demagnetization of the

PMs.

1. Active material cost, AM C, is given before by Eq. 2.2.3.

2. The power losses consist of the loss components introduced in 2.1.1, namely

copper losses in windings, and the stator core losses including hysteresis and

eddy current losses.

3. As mentioned previously, the stator outer diameter and the rotor inner diameter

are held constant. Following the calculation of the average torque per unit

length, using CE-FEA for the rated current density, the stack-length of the

designs are adjusted accordingly to produce the desired average 300 Nm torque

at 1500 rev/min. The torque ripple is subsequently determined from the torque

profile over a full fundamental ac cycle.

4. The degree of PM demagnetization is characterized by the minimum flux density

in the rotor PMs over a complete ac cycle. Demagnetization is considered at

the magnet piece level.

The fitness functions of the optimization problem can now be built upon the

discussed performance metrics. Two different scenarios of objectives and constraints

are pursued for a rigorous comparison between the CMODE and the DE algorithms.

In both scenarios, two objectives subjected to two constraints are considered. The

first scenario consists of the following two objectives and constraints:

• Objectives: (a) Minimization of AMC, and (b) minimization of power losses


100

• Constraints: (a) Torque ripple less than 15% , and (b) PM demagnetization less

than 70% of the retentivity/remnant Br

The second scenario is designated as follows:

• Objectives: (a) Minimization of torque ripple, and (b) minimization of power

losses

• Constraints: (a) Axial stack length less than 200 mm, 70 mm, and 400 mm

for the 48S8P-a, 48S8P-b, and 12S10P machines, respectively, and (b) PM

demagnetization less than 70% of Br

4.4 Comparative Study of the Results

Twelve runs of large-scale design optimizations were carried out on a desktop

workstation using 8 simultaneous processing units and 8 ANSYS Maxwell distributed

solvers. The machines were optimized for the two aforementioned fitness functions,

using either the DE or the CMODE as the stochastic optimizer. The DE consists of 40

generations, each of 80 members. The CMODE starts with an initial generation of 180

members and proceeds with 378 eight-member generations. The number of members

in each generation are recommended by references [77, 96] for DE and reference [151]

for CMODE. The overall number of design evaluations are approximately equal, 3200

designs in DE versus 3204 designs in CMODE.

In Figs. 4.5 through 4.8, the progress of the optimization process in terms of the

conflicting objectives, which are normalized independently for each example machine,

is illustrated for the two sets of fitness functions. The number of the function
101

evaluations, i.e. the sequence of the candidate designs are color coded to provide

an indication of the convergence of the design space to the Pareto front vicinity.

Furthermore, the designs are differentiated based on their feasibility to provide an

indication of the effectiveness of the constraint handling in the two optimization

algorithms. It can be seen in these figures that both DE and CMODE successfully

converge to the same optimal neighborhood in the design space. However, the

concentration of the feasible design candidates in the Pareto front vicinity is larger

for CMODE, resulting in a better-defined Pareto front with a lesser number of design

evaluations as opposed to DE, also see Figs. 4.9, and 4.10. In addition, the color code

in Figs. 4.5 through 4.8 suggests that the convergence to the Pareto front solutions

is faster for CMODE.

A comparison between Figs. 4.5 and 4.6, and Figs. 4.7 and 4.8 reveals that the

realization of the objectives and constraints is more difficult in the second scenario of

the fitness functions. Nonetheless, CMODE is still superior to DE as shown in Figs.

4.7 and 4.8, and in Fig. 4.10.

To further discern the difference between the convergence rates of the two search

algorithms, an auxiliary variable is defined as the normalized product of the two

objectives in the feasible design candidates. The decay of this quantity over simulation

time can serve as an indication of the optimization progress. The mean of this

quantity per each generation of optimization is shown in Figs. 4.11, and 4.12. The

two previous observations regarding the faster convergence rate of CMODE, and the

denser concentration of the feasible design solutions in the vicinity of the Pareto

front, are distinctly verified in these figures. As can be seen in Figs. 4.11, and 4.12
102

4
3000
o feasible
3 × infeasible

FE evaluations
Norm. AMC

2000
2

1000
1
48S8P−a
0 0
0.5 1 1.5 2 2.5
Norm. power loss

2.5
3000
o feasible
2 × infeasible

FE evaluations
Norm. AMC

2000
1.5

1000
1
48S8P−b
0.5 0
0.5 1 1.5 2 2.5 3 3.5
Norm. power loss

5
3000
o feasible
4
× infeasible
FE evaluations
Norm. AMC

3 2000

2
1000
1
12S10P
0 0
0.8 1 1.2 1.4 1.6 1.8 2
Norm. power loss

Figure 4.5: The evolution of the optimization process using DE for the three case-
study motors under the first scenario.
103

4
3000
o feasible
3 × infeasible

FE evaluations
Norm. AMC

2000
2

1000
1
48S8P−a
0 0
0.5 1 1.5 2 2.5
Norm. power loss

2.5
3000
o feasible
2 × infeasible

FE evaluations
Norm. AMC

2000
1.5

1000
1
48S8P−b
0.5 0
0.5 1 1.5 2 2.5 3 3.5
Norm. power loss

5
3000
o feasible
4
× infeasible
FE evaluations
Norm. AMC

3 2000

2
1000
1
12S10P
0 0
0.8 1 1.2 1.4 1.6 1.8 2
Norm. power loss

Figure 4.6: The evolution of the optimization process using CMODE for the three
case-study motors under the first scenario.
104

40
3000
o feasible
30 × infeasible

FE evaluations
Ripple (%)

2000
20

1000
10
48S8P−a
0 0
0.5 1 1.5 2 2.5
Norm. power loss

40
3000
o feasible
30 × infeasible

FE evaluations
Ripple (%)

2000
20

1000
10
48S8P−b
0 0
0.5 1 1.5 2 2.5 3 3.5
Norm. power loss

15
3000
o feasible
× infeasible
FE evaluations

10
Ripple (%)

2000

5 1000

12S10P
0 0
0.8 1 1.2 1.4 1.6 1.8 2
Norm. power loss

Figure 4.7: The evolution of the optimization process using DE for the three case-
study motors under the second scenario.
105

40
3000
o feasible
30 × infeasible

FE evaluations
Ripple (%)

2000
20

1000
10
48S8P−a
0 0
0.5 1 1.5 2 2.5
Norm. power loss

40
3000
o feasible
30 × infeasible

FE evaluations
Ripple (%)

2000
20

1000
10
48S8P−b
0 0
0.5 1 1.5 2 2.5 3 3.5
Norm. power loss

15
3000
o feasible
× infeasible
FE evaluations

10
Ripple (%)

2000

5 1000

12S10P
0 0
0.8 1 1.2 1.4 1.6 1.8 2
Norm. power loss

Figure 4.8: The evolution of the optimization process using CMODE for the three
case-study motors under the second scenario.
106

1.4 DE
CMODE
Norm. AMC

1.2

1
48S8P−a
0.8
0.8 0.9 1 1.1 1.2 1.3
Norm. power loss

1.6
DE
1.4 CMODE
Norm. AMC

1.2

1
48S8P−b
0.8
0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5
Norm. power loss

3
DE
2.5 CMODE
Norm. AMC

1.5

1
12S10P
0.5
0.9 1 1.1 1.2 1.3 1.4 1.5
Norm. power loss

Figure 4.9: Feasible Pareto optimal designs of the two optimization algorithms for
scenario 1 of the fitness functions.
107

25
DE
Ripple (%) 20 CMODE
15

10

5
48S8P−a
0
0.8 1 1.2 1.4 1.6 1.8 2
Norm. power loss

15
DE
CMODE
10
Ripple (%)

48S8P−b
0
0.8 1 1.2 1.4 1.6 1.8 2
Norm. power loss

5
DE
4 CMODE
Ripple (%)

1
12S10P
0
0.9 0.95 1 1.05 1.1 1.15 1.2 1.25 1.3
Norm. power loss

Figure 4.10: Feasible Pareto optimal designs of the two optimization algorithms for
scenario 2 of the fitness functions.
108

the duration and the convergence rate of the two optimization algorithms is very

much dependent on the problem definitions. Yet, compared to DE, CMODE reaches

steady-state at least twice as fast in all the twelve different case studies.

The optimizations can be continued in steady-state until the Pareto front acquires

a well-defined profile. Since the majority of the simulation time is spent to solve the

FE-based models, as opposed to the fraction of seconds spent by the optimization

search algorithms, the total duration of the optimization procedure is equal for the

two algorithms, given that the number of FEs are the same. However, the CMODE

algorithm is able to produce a large number of optimal designs in contrast to the DE

algorithm, for which the simulation needs to be continued.

As shown in Figs. 4.9 and 4.10, and according to Table 4.2, CMODE also provides

a larger number of optimal solutions in the immediate vicinity of the Pareto front,

resulting in CMODE’s higher definition and better diversity of Pareto front solutions.

To quantitatively compare the Pareto fronts of the two optimization algorithms,

the hypervolume indicator [160] is calculated. The hypervolume measures the

dominance of the Pareto front solutions with respect to a reference point in the

design space [161]. Here, it is calculated with respect to the maximum objective

values for each set of Pareto front solutions. The normalized hypervolumes of the

two optimization algorithms in Fig. 4.13 clearly indicates the persistent superiority

of CMODE in terms of diversity and quality of generated Pareto fronts.

The two search algorithms are also compared in terms of their constraint handling

capability in Table 4.3. It can be seen that the overall number of the feasible

design candidates, which pass the two constraints imposed either on the torque ripple
109

2.5
Mean Norm. loss × AMC
DE
CMODE
2

1.5

1 48S8P−a

0 0.8 1.6 2.4 3.2 4


Total time (h)

2.5
Mean Norm. loss × AMC

DE
CMODE
2

1.5

1 48S8P−b

0 0.9 1.8 2.7 3.6 4.5


Total time (h)

3
Mean Norm. loss × AMC

DE
2.5 CMODE

1.5

1 12S10P

0 1 2 3 4 5 6 7
Total time (h)

Figure 4.11: Convergence of the feasible design candidates for the three case-study
motors in terms of loss × AM C for scenario 1.
110

30

Mean Norm. loss × ripple


48S8P−a

20

10
DE
CMODE
0
0 0.8 1.6 2.4 3.2 4
Total time (h)

40
Mean Norm. loss × ripple

DE
30 CMODE

20

10
48S8P−b
0
0 0.9 1.8 2.7 3.6 4.5
Total time (h)

10
Mean Norm. loss × ripple

DE
8 CMODE
6

2
12S10P
0
0 1 2 3 4 5 6 7
Total time (h)

Figure 4.12: Convergence of the feasible design candidates for the three case-study
motors in terms of loss × ripple for scenario 2.
111

DE
1.2
CMODE

Hypervolume (−)
1

0.8

0.6

0.4

0.2

0
48S8P−a 48S8P−b 12S10P
(a) Scenario 1

DE
1.2
CMODE
Hypervolume (−)

0.8

0.6

0.4

0.2

0
48S8P−a 48S8P−b 12S10P
(b) Scenario 2

Figure 4.13: The normalized hypervolumes of the Pareto fronts generated by DE and
CMODE for the two scenarios of fitness functions.

and on the PM demagnetization in the first scenario, or on stack length and PM

demagnetization in the second scenario are higher in CMODE. The effective constraint

handling of the CMODE algorithm, in addition to the denser designs in its Pareto

front vicinity, translates into its superior computational efficiency when compared to

DE.

Typical cross-sections of the optimal design solutions and their field plots for each

scenario are provided in Fig. 4.14.


112

Table 4.2: Comparison of the number of the feasible Pareto optimal designs between
DE and CMODE.
Scenario 1 Scenario 2
Machine
DE CMODE DE CMODE
48S8P-a 17 42 12 28
48S8P-b 24 52 4 10
12S10P 34 41 13 30

Table 4.3: Comparison of the number of total feasible design candidates between DE
and CMODE.
Scenario 1 Scenario 2
Machine
DE CMODE DE CMODE
48S8P-a 1136 1189 1780 2306
48S8P-b 1500 2256 1438 1731
12S10P 3162 3199 1387 1849

Because of the sequential processing of generations in synchronous EAs such as

the above implementations of DE and CMODE, the optimum number of distributed

solvers is equal to the number of members in each generation. It should be emphasized

that in the foregoing comparison between DE and CMODE, eight distributed solvers

corresponding to the number of members in each generation of CMODE were utilized.

This number of parallel solvers can be readily implemented on a single desktop

computer with limited processing power and with a minimum number of software

licenses, e.g. eight processing cores/threads and licenses in this investigation. The

CMODE algorithm will maintain the discussed superiorities in comparison to DE

if a lesser number of distributed solvers were to be used, which is usually the case

for smaller motor design companies/groups. However, for implementation on a large

number of distributed solvers, such as in HPC systems, the speed of simulations in


113

(a) Scenario 1

(b) Scenario 2

Figure 4.14: Typical optimized cross-sections and the field plots of the studied motors
for the two scenarios of fitness functions.

DE as a standard EA will linearly increase up to the number of members in each

generation, whereas the additional computational resources cannot be fully utilized

in the CMODE algorithm as a steady state EA, although CMODE still benefits from

a faster convergence rate per generation and a more effective constraint handling

mechanism. This highlights the importance of choosing the proper optimization

algorithm in reference to the available computational resources, a topic that has

been the focus of this chapter by introducing CMODE for implementation on desktop

computers. Under these premises, the performance of CMODE is expected to surpass

other standard population-based multi-objective EAs which have been compared with

DE in the literature.
114

4.5 Summary

A new CMODE-type algorithm was developed for design optimization of electric

machines with limited computational resources. In the case-study IPM motors

with distinctive stator winding configurations, rotor layouts, and electrical loadings,

CMODE consistently demonstrated a faster convergence, at least twice as fast

as the convergence rate of DE, a higher definition of Pareto front, and a better

constraint handling in comparison with DE. These can be attributed to the distinctive

population evolution model of CMODE, and its effective constraint handling method,

in which the degree of constraint violation is being minimized simultaneously with

other objectives of the optimization problem.

In design problems with more than one constraint, such as the case-study examples

in this chapter, it might be required to introduce particular weights in the summation

operation given in the definition of the degree of constraint violation. These weights

are designated so as to make the violations comparable between all the constraints.

This can be accomplished, for example by normalizing such violations with respect

to their expected range of variations. The assignment of the weights requires expert

knowledge and particular attention in the implementation of CMODE for electric

machinery design problems.

CMODE’s fast convergence and fewer number of simultaneous function

evaluations makes it best suited for implementation in a state-of-the-art

multiprocessor desktop computer with a lesser number of software licenses, as opposed

to the high-end HPC systems.


115

CHAPTER 5
DRIVE-CYCLE PERFORMANCE
OPTIMIZATION

In Chapters 2 and 3, the need for including the entire range of operation in the

optimization algorithm was demonstrated. In this respect, this chapter presents a

large-scale FE model-based design optimization algorithm for improving the drive-

cycle efficiency of PM synchronous machines with wide operating ranges such as

those used in traction propulsion motors. The load operating cycle is efficiently

modeled by using a systematic clustering method to identify the operating points

representing the high-energy-throughput zones in the torque-speed plane. The

machine performance can be evaluated over these cyclic representative points using

the previously introduced CE-FEA approach which is upgraded to include both

constant torque and extended speed regions in the evaluation of the machine

performance metrics. In contrast to the common practice, which aims at enhancing

the rated performance, the entire range of operation is considered in the present

design optimization method. Practical operational constraints imposed by the voltage

and current limits of the motor-drive system, excessive PM demagnetization, and

torque ripple are accounted for during the optimization process. The convergence to

the optimal design solutions is expedited by utilizing the new stochastic optimizer

described in Chapter 4. The developed design algorithm is applicable to any

configuration of sine-wave driven PM and synchronous reluctance motors over any


116

conceivable load profile. The design optimization algorithm is developed in this

Chapter, followed by the case-study investigations in the next chapter.

5.1 Introduction

As was elaborated in the previous chapters, depending on the particular application,

the CE-FEA based design optimization techniques aim at realizing a set of objectives

under certain performance constraints described/embedded in the optimization fitness

function [84, 85, 88, 96, 119, 162]. In line with common practice, the CE-FEA based

methods evaluate the associated fitness function at the rated load point, i.e. base

speed and rated torque without directly incorporating the efficiency requirements of

the extended speed range operation into the optimization fitness function.

One of the pioneering FE-based design optimization efforts for improving the field

weakening performance was introduced in [92] where an IPM motor was optimized

for maximizing electromagnetic torque at base speed and normalized characteristic

current. Similarly, the idea of equality of characteristic current, ICH , with rated

current, IR , was introduced as an additional objective for optimization of IPM motors

for enhanced field weakening performance [95, 96]. In [99], the following objectives

were concurrently pursued: maximization of the torque at base and at maximum

speeds, and minimization of the weight. Although the equality of ICH and IR was not

directly introduced in the fitness function, excluding efficiency from the objectives

resulted in an optimized design in which the two currents were equal [99].

The equality of ICH and IR improves the torque production capability. However,
117

when the non-linear and lossy nature of the machine is considered, from the efficiency

standpoint, congruity of ICH and IR cannot be the ideal criterion for constant power

operation [97, 98]. This will be described in more detail later in this chapter. In [100],

maximizing the torque and efficiency at the base and maximum speeds were pursued

to simultaneously improve the efficiency and operation range of a concentrated flux

IPM motor.

In more recent studies, researchers have attempted to optimize the machine

performance for a specific drive cycle [105, 106, 163, 164]. In [105, 106], a method

known as the cyclic representative points was implemented to efficiently model a

target driving cycle by a finite number of torque versus speed points. These points

were derived based on the energy distribution function specifically calculated for a

given vehicle model and driving cycle. Those investigators subsequently performed

the optimization over these cyclic representative points. However, the selection

criterion of these points requires a more systematic procedure for identification of

the high-energy-throughput zones of the load energy distribution function, especially

when more demanding operating cycles are desired. Furthermore, since a large-scale

optimization was not pursued, the design space was not fully explored.

Most recently, in [107] a propulsion PM assisted synchronous reluctance (SyR)

motor was optimized using the relatively subjective drive cycle modeling method

introduced in [105, 106]. Nonetheless, because of the computationally demanding

nature of the adopted approach, a limited number of design variables were treated,

i.e. tooth width and slot height. Furthermore, the torque ripple was not included in

the optimization process.


118

In this Chapter, the large-scale design optimization of PM machines for a specific

load profile is investigated. The CE-FEA approach is upgraded to enable fast and

high fidelity performance evaluation of the design candidates at any load operating

point residing either in the constant torque or extended speed regions. To further

increase the computational efficiency of the design optimization, the new stochastic

CMODE-type search algorithm presented in Chapter 4 is adopted.

The presented design optimization method is applicable to interior and surface-

mounted PM motors with various slot-pole combinations and rotor magnet layouts,

SyR motors, and in essence any PM motor which, through proper drive controls, is

energized by sinusoidal terminal currents. Moreover, the design optimization can be

performed over any conceivable motor operating cycle, while taking into account the

practical operational constraints imposed by the supply voltage and/or the motor

current limits [165].

Accordingly, in this chapter, the efficient modeling of the motor load profile is

explained first. The control algorithm for derivation of the forcing function for FE

analysis at any load point is described next, followed by the optimization algorithm,

and a discussion on defining appropriate objectives and constraints. Using the

optimization methodologies in this chapter, the optimization of two example traction

motors will be presented in Chapter 6.


119

5.2 Efficient Modeling of the Motor Load Profile

In this section, a systematic method, as an alternative to the method suggested in

[105, 106], is presented for efficient modeling of a given operating cycle using a limited

number of load points.

5.2.1 Energy Distribution Over the Torque-Speed Plane

The first step of the optimization process is the identification of the motor torque

and speed profiles for the specific application, which by itself is a demanding task

and might require extensive data collections, field studies, and system modeling. In

the case of traction motors, which is adopted in this dissertation as a challenging

design example, numerous factors such as the drivetrain technology, transmission

system, energy management unit, driving habits, terrain, and the vehicle operating

mode determine the motor propulsion requirements.

To address the above mentioned complexities in obtaining the motor torque and

speed profiles, here, the Advanced Vehicle Simulator (ADVISOR) developed by the

National Renewable Energy Laboratory (NREL) is used. Two hybrid vehicles, which

were previously discussed in Chapter 2 with regards to the distribution of the load

operating points in their torque-speed plane, are modeled using ADVISOR, namely,

the Toyota Prius Gen. 2, and the Honda Insight Gen.1.

Furthermore, the distinctive propulsion requirements of different driving cycles

are illustrated by modeling these vehicles over two US driving cycles, namely, the

Urban Dynamometer Driving Schedule (UDDS), and the Highway Fuel Economy
120

350 200
300 150

Torque (Nm)
Speed (rad/s)
250 100
200 50
150 0
100 −50
50 −100
0 −150
0 200 400 600 800 1000 1200 1400
time (s)

(a) UDDS

400 200

300 100

Torque (Nm)
Speed (rad/s)

200 0

100 −100

0 −200
0 100 200 300 400 500 600 700 800
time (s)

(b) HWFET

Figure 5.1: The Toyota Prius Gen 2. motor load profiles.

Test Driving Schedule (HWFET). The resulting load profiles of the example traction

motors under the aforementioned two studied driving cycles are shown in Figs. 5.1

and 5.2.

For each vehicle, the wide variations of the motor torque and speed profiles based

on the driving cycles shown in Figs. 5.1 and 5.2 reveal the challenges involved

in designing a motor for efficient and reliable operation at all torque and speed

levels. Furthermore, by comparing the motor load profiles between the two vehicles

in these figures, it becomes obvious that even for this specific traction propulsion

motor application, the preferred motor design and specifications, such as slot-pole
121

500 50
450 40
400 30

Torque (Nm)
Speed (rad/s)
350 20
300 10
250 0
200 −10
150 −20
100 −30
50 −40
0 −50
0 200 400 600 800 1000 1200 1400
time (s)

(a) UDDS

400 20

300 0

Torque (Nm)
Speed (rad/s)

200 −20

100 −40

0 −60
0 100 200 300 400 500 600 700 800
time (s)

(b) HWFET

Figure 5.2: The Honda Insight Gen. 1 motor load profile.

combination, rotor and stator layouts, rated operating conditions, etc. are vehicle-

dependent and one panacea design solution does not exist.

This observation further underscores the need for a systematic design optimization

methodology such as the one developed in this dissertation which can accommodate

the aforementioned factors, and can also provide a simulation platform for evidence-

based comparison of the alternative design solutions.

The motor continuous torque and base speed should be characterized with

reference to the load profiles. The continuous torque from the thermal limit
122

(a) UDDS

(b) HWFET

Figure 5.3: Toyota Prius Gen. 2 motor output energy over torque-speed plane.

standpoint can be determined based on the root mean square of the torque profiles:
s Z t2
1
Trms = T (t)2 dt (5.2.1)
(t2 − t1 ) t1

The determination of the base speed, ωbase , and optimum speed ratio, which

is the ratio of maximum speed, ωmax , to ωbase , depends on many figures of merit

including the system power specifications, total weight, losses, etc. Previous studies

have indicated that this optimum speed ratio falls within a range of 3 to 4 for PM

synchronous motors used in propulsion applications [166, 167].

Using the motor output torque and speed profiles, the absolute value of the
123

(a) UDDS

(b) HWFET

Figure 5.4: Honda Insight Gen. 1 motor output energy over torque-speed plane.

energy distribution over the torque-speed plane, obtained from the torque times speed

product and the corresponding time spent at this condition, can be calculated as

shown in Figs. 5.3 and 5.4.

The motor operating regions, constant torque and constant power, can be

recognized in the distribution of the load points in Figs. 5.3 and 5.4. Once again, the

considerable variations of the load energy distribution functions based on the driving

cycles and the vehicles should be pointed out to the reader.


124

5.2.2 Cyclic Representative Points

Ideally, the motor performance over each individual load point shown in Figs. 5.3

and 5.4 needs to be considered in its drive-cycle performance evaluation. However,

for obvious computational reasons, the swarm of the operating points in the load

energy distribution needs to be modeled by a limited number of so-called cyclic

representative load points which should convey the main features of the driving cycle

in a computationally efficient manner. Specifically, these representative points should

indicate (a) the speed and torque at the high-energy-throughput operating zones of

the torque-speed plane, and (b) the energy weights associated with these zones which

is the measure of significance of these zones in the evaluation of the motor drive-cycle

efficiency.

In [105, 106], the idea of computationally efficient modeling of the energy

distribution by a small number of representative load points was implemented on

an example pure electric vehicle operated over New European (NEDC) and Artemis

driving cycles. In their approach, for characterization of these representative points,

the investigators first partitioned the torque-speed plane according to the locations of

the high-energy-throughput regions. Subsequently, the so-called “centers of gravity”

in these regions were calculated according to Eq. (5.2.3) through Eq. (5.2.4):

Ni
X
Ei = Eij (5.2.2)
j=1
Ni
1 X
ωi = Eij ωij (5.2.3)
Ei j=1
Ni
1 X
Ti = Eij Tij (5.2.4)
Ei j=1
125

where Eij , ωij , and Tij are respectively the energy, angular speed, and torque

associated with each load point in region i, Ni is the number of load points in region

i, and ωi and Ti indicate the center of gravity of the region, which are identified as the

cyclic representative points. The main disadvantage of this approach for identification

of the representative load points is the subjective/observation-based selection of the

regions, as opposed to a mathematical energy minimization-based approach, which

questions the optimality of the discretization process, and also poses an impediment

to its application for more sophisticated energy distribution functions.

Here, a systematic method of quantization that is popular for cluster analysis

in data mining known as the k-means clustering algorithm [168] is introduced for

modeling the load operating cycle. Using this method, all the n observations in the

energy function (x1 , . . . , xn ) can be partitioned into k ≤ n clusters (s1 , . . . , sk ) in

which each observation belongs to the cluster with the nearest mean (m1 , . . . , mk ),

serving as the prototype of the cluster. The standard algorithm used here has two

iterative steps, the assignment step and the update step.

In the first step, each observation, xp , is assigned to the cluster, si , with the

nearest mean, mi , according to Eq. (5.2.5):


(t) (t) (t)
si = {xp |k xp − mi k2 ≤k xp − mj k2 ∀j,
(5.2.5)
1 ≤ j ≤ k}

where t is the iteration count number. The means for the initial iteration, m0i , can

be chosen randomly.

In the update step, the centroids of the observations in the new clusters, mt+1
i ,

given in Eq. (5.2.6) are designated as the new means. These assignment and update
126

Norm. energy (pu)


1

0.5

0
1
1
0.5 0.5
0 0
Norm. torque (pu) Norm. speed (pu)

(a) UDDS
Norm. energy (pu)

0.5

0
1
1
0.5 0.5
0 0
Norm. torque (pu) Norm. speed (pu)

(b) HWFET

Figure 5.5: Cyclic representative points of Toyota Prius Gen. 2 motor output energy
over torque-speed plane.

steps are iteratively repeated until convergence is reached, i.e. until the assignments

of mti in Eq. (5.2.5) do not change.

1 X
mt+1
i = (t)
xj (5.2.6)
| Si | xj ∈S t
i

Using the k-means algorithm, first the normalized energy distribution function

is partitioned into a limited number of clusters, the means of which yield the

representative torque-speed point of that cluster. Thereafter, the energy weights

of each representative point are computed based on the ratio of the average energy
127

Norm. energy (pu)


1

0.5

0
1
1
0.5 0.5
0 0
Norm. torque (pu) Norm. speed (pu)

(a) UDDS
Norm. energy (pu)

0.5

0
1
1
0.5 0.8
0.6
0 0.4
0.2
Norm. torque (pu) Norm. speed (pu)

(b) HWFET

Figure 5.6: Cyclic representative points of Honda Insight Gen. 1 motor output energy
over torque-speed plane.

consumed in the corresponding cluster to the total energy associated with the drive-

cycle.

The cyclic representative points and their associated energy weights for the

previously discussed motors and operating cycles are shown in Figs. 5.5 and 5.6.

It can be seen in these figures that the distribution of the representative load points

and their energy weights vastly vary with respect to the vehicle model and motor

operating cycle, suggesting once again that an optimal design which suits one case
128

might be suboptimal for the other cases.

Although the k-means clustering algorithm is a “computationally difficult”

problem, which requires use of iterative optimization algorithms for obtaining a

solution, it is only run once at the initialization stage of the drive-cycle optimization

to yield the representative load points. The number of clusters can be determined

based on the sum of the distances of the load points to their corresponding cluster

means, as defined in Eq. (5.2.7).

k X
X
Sumof distances = ||xj − mi ||2 (5.2.7)
i=1 xj ∈Sit

This quantity is calculated for the previously discussed vehicles and driving cycles

over a wide range of cluster numbers as shown in Fig. 5.7. For these example

distributions of load points, it appears to this investigator that a “happy balance”

between accuracy and computational speed points to a choice of seven representative

load points corresponding to the seven clusters shown in Figs. 5.5 and 5.6. A larger

number of clusters would provide a marginally more accurate approximation of the

energy distribution function at the cost of diminished computational efficiency, since

the performance evaluation is carried out over every individual cyclic point.

5.3 Performance Evaluation at Representative

Points

After the determination of the cyclic representative points, the motor performance

should be evaluated over them since most of the power is consumed or generated at

these points. Characterization of the motor performance at the cyclic representative


129

Best total sum 80 Toyota Prius Gen. 2 over UDDS


of distances 60
40
20
0
2 4 (a) 6 8 10

30
Toyota Prius Gen. 2 over HWFET
Best total sum
of distances

20

10

0
2 4 (b) 6 8 10

40
Honda Insight Gen. 1 over UDDS
Best total sum
of distances

20

0
2 4 (c) 6 8 10

20
Honda Insight Gen. 1 over HWFET
Best total sum
of distances

10

0
2 4 (d) 6 8 10

Number of clusters

Figure 5.7: The best sum of the distances of the load points to their corresponding
cluster means versus the number of clusters.
130

Norm. torque (pu)


1
Unsaturated
0.5
Saturated
0
1 0
0.5 −0.5
0 −1
iq (pu) id (pu)

(a) Torque

1
Norm. vR (pu)

Unsaturated
0.5
Saturated

1 0
0.5 −0.5
0 −1
iq (pu) id (pu)

(b) Induced voltage

Figure 5.8: Effects of saturation and cross-saturation in prediction of torque, and


induced voltages over the full range of excitation current.

points, which can be located anywhere in the torque-speed plane, requires careful

control of the machine excitation current for production of maximum torque per

ampere (MTPA) under performance constraints imposed either by the motor rated

current in the constant torque region, or by the maximum output voltage of the

supply, rated current, and PM demagnetization in the constant power region.

To characterize the motor performance using high fidelity magnetostatic FEA

simulations, the magnitude and phase angle of the stator winding excitation current
131

needs to be determined for each individual design candidate at each representative

load point. Accurate estimation of the optimum current density and its angle of

advance from the q-axis is imperative in order to ensure a reliable performance

comparison between the design candidates. Here, it should be pointed out that

the linear inductance-based models for IPM machines fail to accurately predict the

machine behavior when saturation and cross-saturation phenomena are prevalent

[163, 169]. This issue is addressed next.

To demonstrate the inaccuracy of the linear-parameter models in predicting the

performance of the high-power-dense PM motors, in Fig. 5.8, the produced average

steady state torque, Tavg , and the fundamental value of the total induced phase

voltage in the stator windings at steady state and under a constant speed, vR , of

the Toyota Prius IPM motor are computed and compared, over the full range of

excitation current for motoring operation, between the linear parameter model given

in Eq. (5.3.1) and Eq. (5.3.2), and the actual values obtained from FEA using Eq.

(5.3.3) and Eq. (5.3.4).

The contributions of the stator winding resistance to the terminal voltage, and the

reduction of torque due to core losses are similar in both cases and are not reflected

in these equations and figures.

3P
Tavg = ((Ld id + λP M )iq − Lq iq id ) (5.3.1)
22
q
vR = ωe ((Ld id + λP M )2 + (Lq iq )2 ) (5.3.2)

3P
Tavg = (λd iq − λq id ) (5.3.3)
22
q
vR = ωe (λ2d + λ2q ) (5.3.4)
132

where P is the number of poles, ωe is the motor speed in electrical rad/s., and Ld and

Lq are respectively the d-axis and q-axis inductance values derived based on three

FEA simulations with different current vectors. That is, the PM flux-linkage, λP M

is obtained at zero current, followed by the calculation of Ld = (λd − λP M )/id and

Lq = λq /iq for small values of excitation current.

It can be seen in Fig. 5.8, that the torque and induced voltage estimation errors

steadily creep up as the current density increases. These errors are more evident along

the q-axis due to higher permeance of the q-path, and due to the demagnetizing effect

of the d-axis component of the armature current. Hence, a new numerical method

with built-in control to conform to the motor-drive system voltage and current ratings

is developed here, see Fig. 5.9.

For each design candidate, CE-FEA with reduced a number of solutions is

performed for various stator excitation currents covering the entire range of motoring

operation. The resulting d- and q-axes flux-linkages, λd and λq , are sampled as the

current vector sweeps through the second quadrant of the d-q plane, and are stored

in corresponding look-up tables. Subsequently, the average value of the steady-state

torque, and the fundamental components of the induced voltage in the stator winding

can be calculated using Eq. (5.3.3) and Eq. (5.3.4), respectively. At this point, the

machine stack length, and the torque and induced voltage look-up tables are scaled

proportionally for production of the required torque at the base speed corresponding

to a current density of Jmax , which is assumed to be the same for all the design

candidates and is determined with reference to the cooling system specifications. The

maximum torque per unit stack length corresponding to the maximum current density
133

Start calculating
current excitations

● Input the design candidate

● Perform CE-FEA with reduced number of solutions


● Obtain torque and flux values at the sampling points
over the current dq-plane
● Generate torque and flux look-up tables
● Scale the design stack length and correspondingly the
torque and flux look-up tables for producing the required
torque at the base speed
● Finely interpolate torque and flux distributions over the
second quadrant of current dq-plane
● Reset the load operating point counter (k = 1)

● Read load operating point mk

● From the torque look-up table, find the set of current


values satisfying the torque at mk , set A
● For each current value in set A, find the corresponding
flux and voltage values from the flux look-up table, set B

● Using set B, find the set of current


values in set A with a corresponding
No voltage less than or equal to the
ω < ωbase rated voltage, set C
● Select the current value with the
Yes minimum magnitude from set C
●k= k+1

● Select the current value with


the minimum magnitude from Yes
k < kmax
set A
●k= k+ 1
No

● Output the set of currents for


all operating points

Stop

Figure 5.9: Developed method for derivation of the stator winding currents at every
load point for time-stepping magneto-static FEA. Optimal control is ensured for
constant torque and flux weakening operation.
134

1 1
1 1

Norm. torque (pu)


0.8 0.8

Norm. λ (pu)
0.8 0.8
0.6 0.6
iq (pu)

iq (pu)

t
0.6 T = 0.5 pu 0.6 ω > ωbase
0.4 0.4
0.4 0.4
0.2 0.2 0.2 0.2

0 0 0 0
−1 −0.5 0 −1 −0.5 0
id (pu) id (pu)

(a) Set of currents which meet the torque (b) Set of currents which meet the voltage
requirements requirements

Figure 5.10: Process of derivation of the excitation current for a typical load point.

in the stator winding, Jmax , indicates the torque production capability of the design

candidate.

After generating the look-up tables and scaling them to produce the required

torque at the base speed, the process of determination of the excitation current can

be started. Instead of fitting a polynomial equation on the torque and induced voltage

samples and using Bisection and Newton-Raphson methods to determine the d- and

q-axes currents for MTPA or field weakening operation as was proposed in [89], the

optimal excitation current for each load operating point is selected through the simple

algorithm shown in Fig. 5.9. Accordingly, for each load point, first the set of current

vectors, id and iq , producing the required torque are identified, set A. The flux values

associated with these current vectors are also stored in set B. If the rotation speed

of the load point is less than the base speed, the vector with the smallest magnitude

from set A is chosen as the optimal current vector for MTPA operation. Otherwise,

if the rotation speed is greater than the base speed, the set of current vectors that do
135

1 1
1 0.8 1 0.8
0.8 0.8

BPM (T)

(T)
0.6 0.6
iq (pu)

iq (pu)
0.6 0.6

PM
0.4 0.4

B
0.4 0.4
0.2 0.2 0.2 0.2

0 0 0 0
−1 −0.5 0 −1 −0.5 0
id (pu) id (pu)

(a) Leading end PM (b) Trailing end PM

Figure 5.11: PM demagnetization maps of a typical motor for PMs located with
respect to the rotor motion for motoring operation.

not violate the constraint on the maximum drive voltage are identified using set B

and equation Eq. (5.3.4). Once again, for efficient operation, the current vector with

minimum magnitude will be selected. This process is illustrated in Fig. 5.10. Using

this method, both MTPA and field weakening operations are successfully incorporated

to ascertain the optimal operation and to maintain current controllability throughout

the extended speed range under limited supply voltage.

Along with sampling the d- and q-axes flux-linkages for calculation of the torque

and induced voltage values, for each individual design candidate, the rotor PM flux

density levels are also obtained by placing “virtual search coils” across the PMs in the

FEA model. The PM flux density levels are used to create the PM demagnetization

maps, such as those shown in Fig. 5.11 for a typical design, in order to characterize

the degree of PM demagnetization throughout the entire possible range of operation.

The minimum PM flux density in the PM demagnetization map should be checked

during the optimization process to prevent excessive/irrevocable demagnetization.


136

5.4 Extended Speed Operation of PM Machines

In the design optimization of high-speed IPM machines, a fundamental step is the

characterization of the machine torque-speed envelope, and investigation of the design

factors which influence the extended speed operation of such machines. Traditionally,

PM motors are categorized into finite and infinite maximum speed classes according

to the ratio of their characteristic current, ICH , with respect to their rated current,

IR [93, 170–172]. A salient pole PM motor that is ideally designed for extended

speed operation is assumed to have ICH equal to IR [93, 96, 170–173]. In [92, 94–96],

this criterion has been introduced as an additional objective for large-scale design

optimization of IPM motors for constant power operation. Yet, such characterization

of an ideal machine for constant power operation falls short to note two factors

regarding realistic operation of such machines. First, in any practical design, the

rotational speed is limited to an upper bound so as to avoid rotor destruction at

excessively high speeds [159]. Second, as will be illustrated here, the equality of

ICH and IR adversely affects the machine’s efficiency throughout the extended speed

range. This is despite the fact that in many applications such as electric vehicles,

efficiency is one of the main concerns.

Another fundamental factor influencing the field-weakening characteristic of a

constant parameter PM motor drive is saliency. As suggested in the literature [93],

increasing saliency can enhance the power capability under the rated voltage and

rated current operating constraints. Nevertheless, increased saliency does not always

translate into reduced cost in IPM motors [96] contrary to a commonly held notion
137

[93]. This can be attributed to the fact that the primary component that contributes

to the saliency of the studied machine configurations in [96] are the permanent

magnets with a recoil permeability close to that of air. Consequently, the saliency in

a particular rotor layout has a direct correlation with the area along the d-axis that

is covered by the expensive magnets. In addition to influencing the cost, saliency

can be conducive to higher efficiencies by contributing to the generation of reluctance

torque from the demagnetizing component of the stator current [93, 170, 171, 174],

particularly in the flux weakening operation. Therefore, it is imperative to assess the

efficiency of the salient pole machines over their entire operating ranges.

Aside from the equality of characteristic and rated currents, (ICH = IR ), which has

been previously suggested for reliable operation [93], and enhanced power capability

throughout the extended speed region, two other possible scenarios for ICH are:

(ICH > IR ) or (ICH < IR ). Provided that (ICH > IR ) as a result of low d-axis

inductance or strong magnets, the extended speed operation will be limited due to

excessive increase of the internal voltage at high speeds. This, in turn, leads to

rapid loss of the control of the stator current, and accordingly a sharp drop of the

generated torque and power. In case (ICH < IR ) due to weaker magnets or higher

d-axis inductance, the infinite speed operation can be achieved at the expense of

diminished output power. The latter case as well as the case in which these currents

are equal (ICH = IR ) are further investigated for the Prius IPM motor configuration.

In Fig. 5.12 the sampled contours of the magnitude of the fundamental component

of the flux-linkage of a typical phase are shown over the entire range of the excitation

current for three IPM motors with equal torque and current density ratings. The
138

1 1
1 0.8 1 0.8

Norm. λR (pu)

Norm. λR (pu)
0.8 0.8
0.6 0.6
iq (pu)

iq (pu)
0.6 0.6
0.4 0.4
0.4 0.4
0.2 0.2 0.2 0.2

0 0 0 0
−1 −0.5 0 −1 −0.5 0
id (pu) id (pu)

(a) Motor 1 (b) Motor 2

1
1 0.8

Norm. λ (pu)
0.8
0.6
iq (pu)

R
0.6
0.4
0.4
0.2 0.2

0 0
−1 −0.5 0
id (pu)

(c) Motor 3

Figure 5.12: Sampled flux contours of the three example motors with equal rated
torque and rated current density.

motors are sorted in descending order of their ICH , which can be identified in these

figures as the point on the negative side of the d-axis where the flux-linkage reaches

zero. To understand how the value of the characteristic current influence the behavior

of the three designs for various torque and speed values, their performance maps are

computed using the FEA-based method described in Fig. 5.9.

The first series of maps in Fig. 5.13 reveals that Motor 1, which has an ICH closer

to its IR , is superior in terms of power capability in the extended speed range as it can

generate more torque with a limited supply voltage. Nevertheless, according to the
139

same figure, the efficiency drops faster for this motor with the increase of the operating

speed. Meanwhile, the designs with lower ICH indicate diminished power capability

in the extended speed region. Yet, the high efficiency contours are further expanded

to cover more areas of the entire operating range in these designs. Accordingly, by

observing Figs. 5.12 and 5.13, one can notice the direct relationship between ICH and

the field weakening power generation capability and the inverse relationship between

ICH and the extended speed operation energy efficiency.

To discern the cause of the differences between the efficiency maps shown in Fig.

5.13, the loss ratio maps over the entire operating range are provided in Fig. 5.14.

The loss ratio is defined here as the ratio of copper loss to the sum of the stator core

loss and the friction and windage loss. The latter is provided by the research team

at Oak Ridge National Laboratories (ORNL) for the 2004 Toyota Prius motor and is

assumed to be equal in all the designs. According to the loss ratio maps, the ratio of

copper loss to core loss is greater in motor 1 in the extended speed range.

The larger high-speed copper losses in motor 1 can be ascribed to the increased

current densities in the extended speed range as shown in Fig. 5.15. Specifically, as

opposed to the comparatively smaller q-axis component of the stator winding current,

Fig. 5.16, a larger negative d-axis component of current, Fig. 5.17, is required to

weaken the magnet flux at high speeds. Accordingly, as ICH increases, canceling

the PM flux-linkage in the stator windings requires larger demagnetizing currents

to either demagnetize the stronger magnets, or to compensate for the lower d-axis

inductance, since by definition ICH = λP M /Ld .

Ultimately, it should be pointed out that although motor 3 has better energy
140

300
95
250

Efficiency (%)
Torque (Nm)
200
90
150
100 85
50
80
1000 2000 3000 4000 5000 6000
Speed (r/min)

(a) Motor 1

300
95
250

Efficiency (%)
Torque (Nm)

200
90
150
100 85
50
80
1000 2000 3000 4000 5000 6000
Speed (r/min)

(b) Motor 2

300
95
250
Efficiency (%)
Torque (Nm)

200
90
150
100 85
50
80
1000 2000 3000 4000 5000 6000
Speed (r/min)

(c) Motor 3

Figure 5.13: Efficiency maps of the three example IPMs with infinite maximum
speeds.
141

300 12
250 10

Torque (Nm)
8

Loss ratio
200
150 6

100 4

50 2
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(a) Motor 1

300 12
250 10
Torque (Nm)

Loss ratio
200
150 6

100 4

50 2
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(b) Motor 2

300 12
250 10
Torque (Nm)

8
Loss ratio

200
150 6

100 4

50 2
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(c) Motor 3

Figure 5.14: Loss ratio maps of the three example IPMs with infinite maximum
speeds.
142

300
15
250

Torque (Nm)

J (A/mm2)
200 10
150
100 5

50
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(a) Motor 1

300
15
250
Torque (Nm)

J (A/mm2)
200 10
150
100 5
50
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(b) Motor 2

300
15
250
Torque (Nm)

J (A/mm2)

200 10
150
100 5

50
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(c) Motor 3

Figure 5.15: Current density maps of the three example IPMs with infinite maximum
speeds.
143

300
15
250

Norm. Jq (A/mm2)
Torque (Nm)
200 10
150
100 5

50
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(a) Motor 1

300
15
250

Norm. Jq (A/mm2)
Torque (Nm)

200 10
150
100 5

50
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(b) Motor 2

300
15
250
Norm. Jq (A/mm2)
Torque (Nm)

200 10
150
100 5

50
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(c) Motor 3

Figure 5.16: Q-axis current density maps of the three example IPMs with infinite
maximum speeds.
144

300
15
250

Norm. Jd (A/mm2)
Torque (Nm)
200 10
150
100 5

50
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(a) Motor 1

300
15
250

Norm. Jd (A/mm2)
Torque (Nm)

200 10
150
100 5

50
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(b) Motor 2

300
15
250
Norm. Jd (A/mm2)
Torque (Nm)

200 10
150
100 5

50
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(c) Motor 3

Figure 5.17: Negative d-axis current density maps of the three example IPMs with
infinite maximum speeds.
145

efficiency in the extended speed region, the magnets are excessively demagnetized for

this design, as shown in Fig. 5.18. This reveals the existing trade-offs between the

ICH , efficiency, power capability, and the degree of PM demagnetization which should

be considered in the design of PM machines for the extended speed operation.

In view of all the above mentioned points, Motor 2 is shown to be an acceptable

compromise among the three motors in terms of power capability, efficiency, and

degree of PM demagnetization. As a matter of fact, this design belongs to the 2004

Toyota Prius IPM motor simulated from the data provided by the team at ORNL. The

following section puts forth a method for designing motors as optimized as the 2004

Toyota Prius IPM motor which takes the aforementioned factors into consideration.

5.5 Drive-Cycle Optimization

5.5.1 Algorithm

After developing the required tools for drive-cycle design optimization, and discussing

the design factors influencing the extended speed performance of PM machines, the

methodology for design optimization of PM machines over a target operating cycle

can be presented through the high-level flowchart of the steps in Fig. 5.19.

The design optimization process has three main stages, namely preprocessing, loop

iterations, and post-processing. In the first stage, the machine model is parametrized

similar to the approach presented in Section 2.2.1, the representative operating points

are identified using the clustering algorithm developed in Section 5.2, the objectives

and constraints of the fitness function are designated considering the discussion in
146

300 1
250 0.8

Min. PM B (T)
Torque (Nm)
200
0.6
150
0.4
100
0.2
50
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(a) Motor 1

300 1
250 0.8

Min. PM B (T)
Torque (Nm)

200
0.6
150
0.4
100
0.2
50
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(b) Motor 2

300 1
250 0.8
Min. PM B (T)
Torque (Nm)

200
0.6
150
0.4
100
0.2
50
0
1000 2000 3000 4000 5000 6000
Speed (r/min)

(c) Motor 3

Figure 5.18: Current density maps of the three example IPMs with infinite maximum
speeds.
147

Start large scale


design optimization

● Input the operating cycle


● Specify the number of
cyclic representative points
● Input the parameterized
machine cross section

● Calculate the load energy function versus torque and


speed to identify the set of all load points
● Implement k-means clustering algorithm on the load
points to obtain the cyclic representative points

● Define the optimization


objectives and constraints
● Initialize the large scale
CMODE-type optimization

● Update the population of the design candidates

● Using the process introduced in Fig. 6, calculate the


current excitation values for the design candidates
● Perform CE-FEA over each cyclic representative point
to calculate performance metrics of each design
● Identify the superior designs

No Stop
criterion
Yes
● Output the Pareto-optimal
design candidates

● Multiphysics performance validation of the selected


design candidates

Stop

Figure 5.19: The flowchart of steps of the overall optimization algorithm.


148

Section 5.4, and the large-scale search algorithm is initialized as explained in Chapter

4. Subsequently, in the loop iteration stage, for each design candidate, the excitation

currents at the cyclic representative points are calculated using the numerical method

developed in Section 5.3. A detailed FE analysis is carried out afterwards over each

individual load point using the fast and high fidelity CE-FEA simulations described

in Section 2.1.1, or full-fledged FE models if desired. Accordingly, the saturation

and cross-saturation are considered both in determining the excitation current and

in calculating the machine performance over each representative load point.

The first iteration of the design candidates is generated randomly with respect

to the designated bounds of the design parameters. The subsequent iterations

are followed by the CMODE-type search algorithm (CMODE). The details of the

CMODE-type optimization and its advantages over the conventional Differential

Evolution have been previously discussed in Chapter 4 for design optimization of

PM motors at the rated operating point. Here, this search algorithm is applied to

optimization of PM machines at multiple operating points.

In each iteration of the CMODE-type optimization process, similar to other

Evolutionary Algorithms (EA), an offspring population competes with the parent

population according a fitness function, i.e. a set of objectives and a set of constraints.

The fitness function is determined based on the design requirements. However,

as illustrated in the previous section, when the non-linear and lossy nature of the

machine is considered, the congruity of ICH , and IR , cannot be the ideal criterion

for constant power operation from the efficiency standpoint. Instead of introducing

such criterion into the optimization fitness function, it is recommended to check the
149

torque production capability of every design candidate at critical load points such as

ωmax , and penalize the designs failing to produce the desired torque under the rated

current and voltage constraints, e.g. by multiplying their associated fitness function

by a large positive number in a minimization problem.

5.5.2 Computational Complexity of the Optimization Algo-

rithm

As discussed in the previous section, the overall optimization process consists of

the preprocessing, loop iteration, and post-processing stages. The computational

burden of the loop iteration stage overshadows that of the other two and is divided

between three subroutines: (a) generation of the torque and flux look-up tables for

calculation of the excitation current, (b) performance evaluation over representative

load points, and (c) determination of the superior design candidates. The first

two subroutines involve FEA, and thus are more computationally demanding. For

generating relatively accurate torque and flux look-up tables, 25 sample current

vectors are recommended, which can be distributed evenly, or can be skewed toward

the negative d-axis to better capture the smaller flux-linkage quantities in this vicinity.

Using CE-FEA, as few as one FE solution can be used for each sample point to extract

the fundamental values of torque and fluxes. Depending on the pole-slot combination,

a larger number of FE solutions are required for calculation of torque ripple and core

losses over each representative load point. Nonetheless, CE-FEA can still be utilized

to significantly expedite the simulation time up to two orders of magnitude when

compared to time-stepping transient FEA [80, 119]. The simulations can be continued
150

until a well defined Pareto front is acquired. Using CMODE-type optimization, this

can be achieved within a smaller number of design evaluations.

5.6 Summary

A novel automated design methodology for optimization of PM synchronous machines

for an application-specific operating cycle was introduced. The developed method

provides a systematic approach for fast and high fidelity design optimization of

PM machines with wide ranges of operation. The constraints imposed either by

the ampere-loading or by the limited drive voltage were fully integrated into the

performance evaluation process, thus enabling the design optimization throughout

the constant-torque and constant-power operating regions. Furthermore, the effects

of magnetic saturation and cross-saturation were thoroughly taken into account both

in determining the current excitation of the stator winding at any load operating

point, and in the calculation of the performance metrics.

Utilizing the k-means clustering algorithm, a systematic method was devised for

efficient modeling of the motor operating cycle. The resultant cyclic representative

points embody the operation zones of the torque-speed plane through which the

majority of the electric energy is consumed. Accordingly, the weighted losses

are derived and incorporated in assessing the drive-cycle efficiency of the design

candidates in the introduced optimization algorithm.


151

CHAPTER 6
CASE STUDY DRIVE-CYCLE OPTIMIZATION
OF TRACTION MOTORS

In this chapter, the optimization algorithm developed in Chapter 5 is utilized

for optimizing two case study PM machines. Each case includes multiple steps,

namely, the modeling of the motor operating cycle, development of the parametrized

FE model, definition of the fitness function, initialization of the search algorithm,

performance evaluation over representative load points, and identification of the

superior designs. A rigorous post-optimization modeling is also pursued to accurately

characterize the performance of the final selected designs.

6.1 Introduction

Two traction propulsion motors are optimized using the techniques presented in the

previous chapters:

• The first motor studied in this chapter is the well-established Toyota Prius

Gen. 2 IPM motor configuration which was previously investigated in a

design optimization for improving its nominal (peak) performance in Chapter

2. This specific motor configuration is chosen here for a twofold purpose: (a)

the particular features of the V-type IPM motors with distributed windings

including saliency and mechanical robustness, which makes them attractive for

high-speed operation; and (b) for performing a detailed comparison between the
152

optimization results and the experimental verifications documented in several

reports published by the Oak Ridge National Laboratory (ORNL) research team

during the past years [130, 131, 175].

• The second case study presented in this chapter is the design optimization of

a spoke-type IPM with a fractional slot concentrated winding configuration

for a formula E racing car propulsion application. The original motor, which

has been successfully tested by other investigators in [142], features a very

high torque per weight ratio. Here, this motor is optimized to further increase

the power density and, at the same time, the drive-cycle efficiency in order to

demonstrate the effectiveness of the developed methodologies in a challenging

design problem. The results of this case study optimization are also utilized to

quantify the performance trade-offs for increasing the power density in spoke-

type PM motors. These trade-offs include the impacts on other performance

metrics such as power losses, PM demagnetization, and torque ripple.

6.2 Optimization of the Prius IPM Motor Over a


Compound Operating Cycle
6.2.1 Initialization

In this section, the 48-slot 8-pole IPM motor of the Toyota Prius Gen. 2 is optimized

over a sequence of driving cycles composed of the US Environmental Protection

Agency’s (EPA) Urban Dynamometer Driving Schedule (UDDS), Supplemental

Federal Test Procedure Driving Schedule (US06), Highway Fuel Economy Test

Driving Schedule (HWFET), Unified Dynamometer Driving Schedule (LA92), and


153

400

Torque (Nm)
200

−200
0 1000 2000 3000 4000 5000
time (s)
(a) Output torque

600
Speed (rad/s)

400

200

0
0 1000 2000 3000 4000 5000
time (s)
(b) Speed

Figure 6.1: Load profile of the Toyota Prius Gen. 2 IPM over a combined driving
cycle.

UDDS. The motor torque and speed profiles which are obtained using ADVISOR are

shown in Fig. 6.1. This sequence of driving cycles are chosen so as to improve the

motors’ performance over a wider variety of operating conditions.

From the load profile provided in Fig. 6.1, the energy distribution versus the

torque and speed can be calculated as shown in Fig. 6.2.

Using the k-means clustering algorithm introduced in Chapter 5, the energy

distribution function is modeled by seven clusters as can be seen in Fig. 6.3. These
154

(a) Isometric view

400 100

300 80
Torque (Nm)

Energy (kJ)
60
200
40
100
20
0
0 100 200 300 400
Speed (rad/s)
(b) Bird’s-eye view

Figure 6.2: Toyota Prius Gen. 2 motor output energy versus torque and speed for
the combined US driving cycles.

load points which will be used in the evaluation of the motor performance during

the optimization process, along with two critical load points imposed by the design

requirements, i.e. required torque at base and maximum speeds are listed in Table

6.1. The numbers next to the means of the clusters in Fig. 6.3 indicate their ranks

according to their energy weights in Table 6.1.

The next step in the drive-cycle design optimization is the definition of the

parametric FE-model, which for this case study has been previously presented in
155

Norm. energy (pu)


1

0.5

0 6
1 2 5
4 1 3 1
0.5 7
0.5
0 0
Norm. torque (pu) Norm. speed (pu)
(a) Isometric view

1
Norm. torque (pu)

0.8
6
0.6

0.4
4
0.2
7 1 2 5
3
0
0 0.2 0.4 0.6 0.8 1
Norm. speed (pu)
(b) Bird’s-eye view

Figure 6.3: Cyclic representative points with seven clusters for the combined US
driving cycle.

section 2.2. This model with the same design parameters and boundaries are used

here. Similar to the original Toyota Prius motor design, rated stator winding current

density is considered to be 16A/mm2 for producing a torque of 300 Nm.

In line with the discussion about proper definition of the optimization fitness

function, which was presented in Section 5.4, the optimization objectives are

designated as follows:

• Minimization of the active material cost given in Eq. (2.2.3);


156

Table 6.1: Cyclic representative points for the combined US driving cycle shown in
Fig. 6.1.
mi ω(rad/sec) T (N m) energyweight
ωbase 157 300
ωmax 628 50
1 140 21 0.2570
2 311 15 0.2273
3 212 19 0.1366
4 74 123 0.1338
5 403 15 0.1265
6 151 288 0.0687
7 44 36 0.0501

• Minimization of the aggregate weighted loss per output power, Pw , defined in

Eq. (6.2.1);
X
Pw = (Pdc,i + PF e,i ) · wi /(Ti ωi ) (6.2.1)
i

where wi is the energy weight of the ith representative load point in Table 6.1,

and ωi is the rotor speed in mech.rad./sec.

Furthermore, the following constraints are imposed to ensure reliable performance

throughout the entire range of operation:

• Less than 25% torque ripple at the rated operating point;

• Less than 70% PM demagnetization at any point on the PM demagnetization

maps.

The thermal aspect of the design is indirectly addressed in the optimization process

by confining the highest current density in the stator winding to that reported for

the Toyota Prius IPM motor. In general, the optimized designs are expected to be

more efficient than the original design, ensuring that the cooling system can properly
157

conduct the power losses to the ambient surroundings. Still, here in this work, the

thermal performance of the most promising design candidates are investigated over a

rigorous driving cycle in a post-optimization stage.

From the mechanical design standpoint, the thickness of the rotor bridges is

adjusted as required to withstand the maximum tangential stress acting on these

bridges. This adjustment is done using approximate calculations of the centrifugal

forces based on the material properties and the shape on the rotor pole-pieces. The

method for analytical estimation of the centrifugal forces is presented in Chapter

7. Minimum thickness is desired for efficient utilization of the magnet flux linkage

and optimal electromagnetic performance, whereas a higher thickness provides better

structural robustness for high-speed operation. A post-optimization mechanical FE

analysis is conducted on the selected optimized designs to make sure they pass this

criterion.

6.2.2 Optimization Results

The drive-cycle design optimization of the 48-slot 8-pole IPM machine with

the aforementioned fitness function was carried out over 10,000 designs using 8

simultaneous processing units on a desktop workstation. The global minimum of

the design space, and accordingly, optimal design solutions were identified within the

first few hundred design evaluations, in this case in less than 24 hours. However,

the optimization iterations were continued to capture a very detailed Pareto front.

The performance of the design solutions which pass the constraints on the PM

demagnetization and torque ripple are shown in Fig. 6.4(a) and (b).
158

1.3

Weighted loss (pu)


1.2

(T)
0.5

PM, min
1.1
0.4

B
1 P
D1
D2
0.9 0.3
0.8 1 1.2 1.4 1.6 1.8 2
Active material cost (pu)
(a) Color-coded for minimum flux density of the rotor PMs

1.3 25
Weighted loss (pu)

Torque ripple (%)


20
1.2
15
1.1
10
1 P
D1 5
D2
0.9 0
0.8 1 1.2 1.4 1.6 1.8 2
Active material cost (pu)
(b) Color-coded for torque ripple

Figure 6.4: Optimization results of Toyota Prius Gen .2 IPM motor over the combined
US driving cycles.

Using the same simulation methodology, the Toyota Prius IPM motor drive-cycle

performance, denoted by P, is also evaluated and marked along with other results

in Fig. 6.4. It can be seen that the Prius design is adjacent to the Pareto-optimal

designs at the high-loss, low-cost vicinity.

The ability of the developed design optimization package in providing design

solutions comparable to the Prius design should be noted. Furthermore, there

are other alternative designs that, at a slightly higher cost, demonstrate better
159

(a) Cross-sections

D1 P D2

111.9

110.5
106.8

105.0

104.3
104.1

102.6

102.1
101.5

101.3
100.1
99.5

99.2
97.5
96.0
100

100

100

100

100

100

100

100

100

100
93.3
NORM. TO P

89.0

87.4

80.3
76.8
AMC T O T A L LO S S LO S S C U LO S S F E EMF THD T R IP P LE M IN . P M B MASS PM MASS CU MASS FE

(b) Electromagnetic performance metrics normalized to Prius

Figure 6.5: Comparison of the selected designs obtained from drive-cycle


optimization.

performance in terms of aggregate power losses. Two of these design candidates

specified by D1 and D2 in Fig. 6.4 are selected for further multi-physics investigation.

6.2.2.1 Electromagnetic Performance

The cross-sections of the three design candidates, D1, P, and D2, and their

performance metrics, normalized with respect to the Prius motor design, are shown

in Figs. 6.5. In Fig. 6.5 (b), the three loss components, are the sum of the respective

losses over the representative load points weighted by their associated energy weights.

Furthermore, the Total Harmonic Distortion (THD) of the induced Electromotive

Force (EMF), and the torque ripple are considered at the rated load point.

The efficiency maps of the three designs is computed by FE analysis of 1 600 sample

load points equidistantly distributed throughout the torque-speed plane, see Fig. 6.6.

The excitation current at each sample point is considered under optimal voltage and
160

300
95
250

Efficiency (%)
Torque (Nm) 200
90
150
100 85
50
80
1000 2000 3000 4000 5000 6000
Speed (r/min)
(a) D1

300
95
250

Efficiency (%)
Torque (Nm)

200
90
150
100 85
50
80
1000 2000 3000 4000 5000 6000
Speed (r/min)
(b) P

300
95
250
Efficiency (%)
Torque (Nm)

200
90
150
100 85
50
80
1000 2000 3000 4000 5000 6000
Speed (r/min)
(c) D2

Figure 6.6: Efficiency maps of the selected designs obtained from drive-cycle
optimization.
161

Figure 6.7: Tested efficiency map of the Toyota Prius Gen. 2 IPM motor reported
by the research team at ORNL.

current control using the method developed in Section 5.3(B). The efficiency maps

should be examined in two specific aspects: first, the highest achievable efficiency,

and second, the extended range of high efficiency contours. The latter is of significant

importance for motor designs in traction applications, and, in general, for applications

in which the motor is to be operated at various load operating points.

To verify the validity of the simulations, the experimentally obtained efficiency

map of the Prius motor reported by the ORNL research team is presented in

Fig. 6.7. The slight discrepancy between the efficiency maps in Figs. 6.7 and

Fig. 6.6(b) can be attributed to the loss components that were not addressed

in our 2-D FE calculations including rotor core losses, eddy current losses in the

magnets, unaccounted temperature variations, and AC conductor losses. Moreover,

the excitation current was assumed to be sinusoidal over the entire operating region

ignoring the time harmonics introduced by the inverter pulse width modulation

specifically in the field weakening region. Nonetheless, the simulation results show

very close correlation to the experimental results. The correlation of the three designs
162

Figure 6.8: Lumped thermal network model of the motor cooling system developed
in Motor-CAD.

not only indicates the merits of the original Prius motor design but also confirms the

effectiveness of the developed design optimization algorithm.

6.2.2.2 Thermal Performance

To compare the thermal performance of the counterpart design solutions relative to

each other, a typical liquid-based cooling system with oil-forced convection through

the housing water-jacket was developed in Motor-CAD. The analysis is based on a

lumped thermal network model as shown in Fig. 6.8 . The coolant fluid is ethylene

glycol compound with 0.375 W/m.C thermal conductivity, 1045 kg/m3 density, and

0.0008987 kg/m/s dynamic viscosity. A constant volume flow rate of 9 l/min at an

inlet temperature of 105 ◦ C was considered for the fluid.

The temperatures of various motor components were obtained by performing a


163

140
135

Temperature (oC)
130
125 P-Winding
120 D1-Winding
115 D2-Winding
110
105
100
0 2000 4000 6000 8000 10000 12000
time (s)
(a) Temperature of stator windings

135
130
Temperature (oC)

125 P-Magnet
120 D1-Magnet
115 D2-Magnet
110
105
100
0 2000 4000 6000 8000 10000 12000
time (s)
(b) Temperature of rotor PMs

112
110
Temperature (oC)

108
P-Housing
106 D1-Housing
D2-Housing
104
102
100
0 2000 4000 6000 8000 10000 12000
time (s)
(c) Temperature of active housings

Figure 6.9: The peak temperatures of the counterpart designs evaluated over US06
driving cycle.
164

transient thermal analysis over the rigorous US06 driving cycle which is characterized

by frequent acceleration and deceleration at various torque and speed levels. It can

be seen in Fig. 6.9 that with identical cooling systems, the temperatures of the stator

windings, rotor PMs, housings, and bearings in the P and D1 counterpart designs

closely correspond. These temperatures are slightly lower in the D2 design due to

the higher efficiency of this design over a broader range of operating conditions as

illustrated in Fig. 6.6(c). The lower operating temperatures can extend the life-time

of motor D2, and thus justify the increased material cost of this design.

6.2.2.3 Stress Analysis of the Rotor Bridges

The mechanical stresses on the rotor bridges are mainly due to the centrifugal forces

resulting from cavities housing the PMs in the rotor structure [159]. A detailed static

structural FE analysis is carried out in ANSYS under steady state maximum speed

of 6 000r/min. It is assumed that forces of electromagnetic, vibration, and rotor

dynamic origins are negligible. Furthermore, it is assumed that the rotor PMs are

not bonded to the cavities since the bonding strength is not permanently constant

and diminishes over time. In the analysis, the mass densities of the rotor laminations

and NdFeB magnets are 7850 kg/m3 , and 7500 kg/m3 , respectively. As can be seen in

Fig. 6.10, the results of the structural analysis demonstrate that the von-Mises [176]

stress throughout the rotor structures of the selected optimized D1 and D2 designs

are comparable to that of the original P design, and are less than the yield strength

of laminations, which is 350 M P a.


165

(a) D1 (b) P

(c) D2

Figure 6.10: Von-Mises stress throughout the rotor structure of the counterpart
designs.
166

6.3 Optimization of a Formula E Racing Car IPM


Motor Over the Le Mans Operating Cycle

In this section, the design optimization of an IPM motor with very high power

density is investigated. Specifically, the optimization of a concentrated flux spoke-

type permanent magnet (PM) motor is in order. These motors can be designed for

increased power density using high-cost high-energy PM materials [177, 178], or can

be designed for reduced dependency on rare-earth materials using alternative low-cost

magnets [179–181]. This is due to the higher air-gap flux density, Bg , of spoke-type

PM motors as given in Eq. (6.3.1) which by itself is in part because of the lower

reluctance of the rotor flux path [177, 178].

−1
h0g

Bg π 2dr
= + 2µmr (6.3.1)
Br 4kσ P hpm wpm

where Br is the PM remnant flux density (retentivity), dr is the rotor outer diameter,

P is the number of poles, hpm is the PM height along the radius of the motor, kσ is

the rotor leakage coefficient, µmr is the PM relative permeability, h0g is the air-gap

height adjusted to account for saturation and slotting effects, and wpm is the PM

width along the magnetization direction.

There are two aspects in the study of high power density motors which require

special treatment, namely, the nonlinearity of the ferrous core, and the thermal

behavior of the machine. Equation (6.3.1) has been used for assessment and rough

approximation of high power density spoke-type machines in many studies such as in

[179, 180]. However, as reported in [179], the effects of saturation and slotting cannot

be accurately modeled in such machines using Eq. (6.3.1), due to the excessive
167

saturation levels of the ferrous core. This fact leaves numerical methods as the

only means for accurate investigation of the design traits in such high power density

machines. Meanwhile, the power density cannot be the sole objective of any realistic

design practice without simultaneously considering the generated power losses, which

directly impact the thermal performance of the machine as the ultimate limiting

factor in increasing the power density. Therefore, a multi-objective approach should

be pursued for the design of such high power density motors.

In this section, an 18-slot, 16-pole spoke-type PM motor configuration is adopted

in order to achieve high drive-cycle energy efficiency, and high torque density for

a direct drive racing car application. Several initial design measures are assumed

in order to realize these objectives, including: (a) the use of special materials

for lamination steels and PMs, i.e. non-oriented thin-gage laminated steel, and

thermally robust SmCo magnets, (b) the adoption of special construction methods

for minimizing PM and winding eddy current losses, i.e. PM segmentation and

twisted wires, and (c) the utilization of a highly efficient cooling system with forced

oil convection through the slot and forced air convection in the air-gap.

Through the FE-based multi-objective large scale design optimization process

developed in the previous chapters, increase of torque ratio per weight (TRW), and at

the same time decrease of power losses over the entire motor load profile are pursued.

The optimization results lay the ground for an FE-based study of high power density

spoke-type motor design in a practical approach, as has been the goal of many previous

investigations using semi-quantitative analytical/closed-form methods [179, 180]. The

existing performance trade-offs in terms of losses, PM demagnetization, and torque


168

100

Torque (Nm)
50

0
0 10 20 30 40 50 60 70
time (s)

(a) Output torque

1500
Speed (rad/s)

1000

500

0
0 10 20 30 40 50 60 70
time (s)

(b) Speed

Figure 6.11: Formula E motor load profile for the Le Mans driving cycle.

ripple for achieving high power density are quantified, and the relationships between

optimal design parameters for high drive cycle efficiency and maximum torque density

are examined. Furthermore, an optimized design which establishes a higher torque

density than the original Formula E racing car motor is achieved.

6.3.1 Initialization

The motor load profiles over the Le Mans driving cycle are shown in Fig. 6.11. The

frequent and oftentimes large fluctuations of the torque profile in Fig. 6.11(a), and

the high speed of operation in Fig. 6.11(b) are characteristics of an electric propulsion
169

(a) Isometric view

100 8

80
Torque (Nm)

Energy (kJ)
60
4
40
2
20

0 0
0 200 400 600 800 1000 1200
Speed (rad/s)
(b) Bird’s-eye view

Figure 6.12: Formula E motor energy consumption in the torque-speed plane.

motor in such a racing car application, which underscore the challenges involved in

the design of such a motor-drive system and its cooling apparatus.

Using the motor load profile, the distribution of the motor output energy consumed

over the entire driving cycle can be obtained from the torque times speed product

and the corresponding time spent at this condition, as shown in Fig. 6.12. In Fig.

6.12(b), the concentration of the load operating points along the maximum torque

line reveals the importance of efficient and reliable field weakening operation of this

motor.
170

Norm. energy (pu)


1

0.5

0
1
1
0.5 0.5
0 0
Norm. torque (pu) Norm. speed (pu)

(a) Isometric view

1
Norm. torque (pu)

0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1
Norm. speed (pu)

(b) Bird’s-eye view

Figure 6.13: Le Mans driving cycle representative points obtained from k-means
algorithm with seven clusters.

The high energy-throughput-zones in the torque-speed plane are identified through

the k-means clustering algorithm with seven clusters. The number of the clusters is

chosen in order to, on the one hand, provide a reasonable estimation of the energy

distribution function, and on the other hand, efficiently model the multitude of the

load operating points with a small number of cyclic representative points.

The centroids of these clusters are listed in Table 6.2 and are marked in Fig. 6.13.

The centroids of the top five clusters, over which most of the energy is consumed, are
171

Table 6.2: Cyclic representative points of Le Mans driving cycle shown in Fig. 6.13.
i ω (rad/sec) T (N m) energy weight w
ωbase 628 105 NA
ωmax 1257 40 NA
1 814 61 0.2245
2 1184 42 0.2213
3 995 49 0.2161
4 631 77 0.1675
5 458 97 0.1343
6 348 63 0.0298
7 409 16 0.0065

used for drive-cycle performance evaluation of the design members throughout the

optimization process.

It is interesting to notice that, similar to the distribution of the load operating

points, the majority of the cyclic representative points are located on the envelope of

the torque-speed plane in the extended speed range in this racing drive cycle.

A parametric 2-D FE model of the all-tooth-wound machine is developed in Ansys

Maxwell for the 18-slot, 16-pole spoke-type IPM motor. The parametrization of the

stator open slot structure and the rotor spoke-type layout is described in Appendix I.

As can be seen in Fig. 6.14, the SmCo PMs are segmented radially as well as axially

to minimize eddy current losses in the magnets. The designated design variables and

their bounds are listed in Table 6.3. They are so defined to allow full exploration of the

design space without causing any interference between surfaces of various components.

The main geometric design constraints are the stator outer diameter and the rotor

inner/shaft diameter which are equal to 80mm, and 30mm, respectively. Accordingly,

the stack length of each design candidate is adjusted to produce a torque of 110 N m

at 6,000 r/min. Furthermore, for all the design members throughout the optimization
172

hy
rso
Stator
Slot
ds wt

rsi hg
dbr wbr

Rotor
hpm
wpm
PM

rri

Figure 6.14: The parameterized FE model of the spoke-type PM motor, see Table
6.3.

Table 6.3: Independent design variables and their upper and lower bounds of the
18-slot 16-pole spoke-type machine.
Parameter(xi ) Description xi,min xi,max
ksi rsi /rso 0.60 0.75
hg Fig. 6.14 0.7 mm 2.5 mm
kwt wt /αs 0.45 0.75
khpm hpm /hpm,max 0.55 0.95
kwpm wpm /wpm,max 0.2 0.6
kwbr wbr /wpm 0.35 0.65
dbr Fig. 6.14 1.5 mm 3.0 mm
hy Fig. 6.14 7 mm 15 mm
173

process, the slot fill factor is assumed to be 0.4, and the temperatures of the stator

windings and PMs are assumed to be 160 ◦ C, and 120 ◦ C, respectively.

Due to heavy saturation and cross-saturation of the high power density motor,

linear parameter models cannot be used to characterize the performance of the design

candidates. Instead, the two-stage approach presented in Section 5.3 is adopted

for this purpose. In the first place, FE simulations with a minimum number of

magnetostatic solutions are used to generate look-up tables of samples of the stator

winding d-axis and q-axis flux linkages over the second quadrant of the current dq-

plane for motoring operation. Using the flux linkage look-up tables, the developed

torque and the induced voltages are computed, and are subsequently incorporated for

determination of the current excitation required for producing the torque and speed

at each representative operating point. The current excitation is determined taking

into account the current and voltage limits of the motor-drive system in the constant

torque, and field weakening regions.

Upon derivation of the current excitation, a detailed FE solution is carried out at

each representative load point to calculate: (a) power losses including the core losses,

copper losses, and the PM eddy current losses, (b) torque ripple, and (c) minimum

flux density in the PMs. As for the latter, the PM demagnetization is also evaluated

under short circuit conditions when the rated current is fully imposed on the negative

d-axis direction. These performance metrics, in addition to the machine total mass,

are subsequently used to compare the merits of the different design solutions in the

optimization algorithm.

After obtaining the machine performance metrics over each operating point, the
174

60 0.5

0.4

(T)
40

Mass (kg)
0.3

PM, min
0.2
20

B
D1
O 0.1
D2
0 0
0 50 100 150
Weighted loss (pu)

(a) Color-coded for PM minimum flux density

60 10

Torque ripple (%)


40
Mass (kg)

4
20
D1
O 2
D2
0 0
0 50 100 150
Weighted loss (pu)

(b) Color-coded for torque ripple

Figure 6.15: The results of optimization of the spoke-type IPM over 3,400 design
solutions.

CMODE-type search algorithm is utilized for converging toward the optimal design

solutions. In this multi-objective design optimization, the two following objectives

are considered:

1. Minimization of the machine weight for the given peak torque of 105 N m at

6 000r/min.

2. Minimization of losses over the high-energy-throughput zones according to Eq.

(6.2.1).
175

60
Pareto−front designs

40

Mass (kg)
20

0
0 50 100 150
Weighted loss (pu)

(a) Selected Pareto-optimal designs

1
Mass
Correlation factors

0.5 Loss

−0.5

−1
m P T B m m
total loss ripple PM,min PM Cu

(b) Correlation factors

Figure 6.16: Correlation of performance metrics with mass and power losses in the
designs optimized for efficiency and high power density.

In addition, two performance constraints are defined in order to:

1. Restrain the maximum toque ripple below 5% at all operating points.

2. Prevent the excessive demagnetization of the rotor PMs over the entire range

of current excitation, i.e. to maintain the degree of PM demagnetization above

20% of the retentivity.


176

6.3.2 Optimization Results

The drive-cycle design optimization of the spoke-type IPM motor over the Le Mans

cycle under the aforementioned objectives and constraints was carried out for 3 400

design candidates. The results of this large-scale design optimization shown in Fig.

6.15 suggest that motor mass is strongly related to other design characteristics which

is described in the following section.

6.3.2.1 Performance Trade-Offs for Achieving High Power Density

To investigate these relationships, Pearson correlation coefficients given in Eq. (6.3.2)

were calculated for the 100 Pareto-optimal designs marked in red in Fig. 6.16(a).

Cov(X, Y )
ρX,Y = (6.3.2)
σX σY

where X and Y are the statistic population of the respective design characteristics

in the 100 Pareto optimal members, Cov is the covariance, and σ is the standard

deviation.

Accordingly, as can be seen in Fig. 6.16(b), the total masses of the selected designs

are inversely proportional to power losses, torque ripple, and mass of PM and copper.

In other words, to increase the power density, larger amounts of PM and copper are

to be used. Meanwhile, increased power density translates to higher power losses,

larger torque ripple, and greater demagnetization of the PMs.

It is interesting to note that as opposed to the torque density, the power

losses constitute a strong positive correlation with the copper mass, suggesting the

conflicting influence of copper mass on the two objectives in a multi-objective design

approach. Furthermore, the negative correlation of power losses with PM mass


177

x
i,max

xi,mid

x
i,min
ksi hg kw kh kw kw dbr hy
t pm pm br

(a) Box plots of the design paramteres

1
Mass
Loss
Correlation factors

0.5

−0.5

−1
ksi hg kw kh kw kw dbr hy
t pm pm br

(b) Correlation factors

Figure 6.17: Distribution of the design parameters in the Pareto-optimal designs, and
their correlation with total mass and power losses.

indicates that the designs with greater power losses are expected to utilize less PM

material.

To unscramble the relationships between the design parameters listed in Table 6.3

and the power density and power losses, the distribution of the design parameters

within their predefined bounds in the Pareto-optimal designs, Fig. 6.17(a), and their

correlations with the two objectives, i.e. power density and drive-cycle efficiency, Fig.

6.17(b), should be examined.


178

rso
Stator
rsi μFe =∞

hg
Rotor
μFe =∞ PM

hpm wpm

τpm

Figure 6.18: Equivalent machine model used for analytical investigation of the optimal
design of spoke-type FSCWs.

Accordingly, the design parameters which constitute negative correlations with

both mass and losses, and thus are relatively confined within narrower upper bands are

the split ratio, ksi , and the tooth-stem width ratio, kwt . In essence, to simultaneously

achieve high power density and high efficiency, these two parameters are expected to

assume higher values.

Base on an analytical investigation of the optimal design of spoke-type FSCWs,

with assumptions of infinite core permeability and slot-less stator as illustrated in

Fig. 6.18, it was reported in [179] that the optimal values for the two critical

design parameters, namely, the split ratio, ksi = rsi /rso , and PM thickness ratio,

kwpm = wpm /τp , are respectively in the range of 0.5-0.55, and 0.4-0.5, respectively, for

a high power density 18-slot 16-pole machine, similar to the configuration studied in

this paper. However, according to the results of the large-scale design optimization,
179

Histogram of ksi
4

0
0.6 0.65 0.7 0.75 0.8
Split ratio
(a)

8
pm

6
Histogram of kw

0
0.2 0.3 0.4 0.5 0.6 0.7 0.8
PM thickness ratio
(b)

Figure 6.19: Histogram of the distributions of (a) ksi , and (b) kwpm in the 100 Pareto-
optimal designs.

in a practical case, when non-linearity of the ferrous core, various loss components,

and complex geometry of the machine is fully taken into account, optimal ranges

other than those derived by analytical models are obtained, see Fig. 6.19 for the

histogram of the ksi , and kwpm parameters in the Pareto-optimal designs. These results

underscore the importance of using high-fidelity models for derivation of optimal

design rules in such high power density and highly saturated machines.
180

Table 6.4: The design characteristics of the counterpart spoke-type motors.


mtotal (kg) Ploss (pu) Tripple,rated (%) BP M,min (T ) mP M (kg) mCu (kg)
D1 16.1 52.0 3 0.31 2.14 0.7
O 9.1 71.7 4.5 0.29 1.00 0.7
D2 5.9 81.6 4 0.36 0.98 0.8

6.3.2.2 Increasing Efficiency or Power Density

From the Pareto-optimal design solutions, two feasible design candidates, which do

not violate the performance constraints imposed on the torque ripple and degree of

PM demagnetization, denoted by “D1”, and “D2” in Fig. 6.15, are selected. The

designs “D1”, and “D2” feature minimum gross power losses, and maximum power

density, respectively . A comparative analysis between these designs with respect to

a design denoted by “O” in Fig. 6.15, which establishes a compromise performance,

is carried out in this section.

The cross-sections and the flux line distributions under rated load conditions of

these three machines are shown in Fig. 6.20. Their main design features are compared

in Table 6.4. As can be seen in Fig. 6.20, the saturation becomes more prevalent

throughout the ferrous core as the power density increases.

The efficiency maps of the three counterpart designs are provided in Figs. 6.21(a)

through (c). The high efficiency contours of the optimized design “D1” are expanded

towards the representative load points with the highest energy weights, see Table 6.2,

which are distributed along the high torque, and high speed vicinity. The design

“D2” has diminished efficiency, mainly due to the elevated copper losses, as opposed

to the core losses which are confined by the saturation phenomena and by the lower
181

(a) D1 (b) O

(c) D2

Figure 6.20: Flux lines distributions of the optimized designs at 6 000 r/min under
rated load.
182

Table 6.5: The design characteristics of the optimized high power density “D3” motor.
mtotal (kg) Ploss (pu) Tripple (%) BP M,min (T ) mP M (kg) mCu (kg)
Original Formula E (AIM) 9.1 71.7 4.5 0.29 1.0 0.7
Optimized Design 7.7 70.6 3.3 0.34 1.0 0.7

volume of the magnetic core in this design. This can be seen in the copper loss and

core loss maps provided in Figs. 6.22 and 6.23, respectively.

6.3.2.3 Increasing Efficiency and Power Density

The results of the optimization suggest that although the original design is a very well

conceived one, its mass and gross losses can be further decreased simultaneously. From

the Pareto-optimal design solutions, a design candidate denoted by “D3” in Fig. 6.24,

which features a higher power density and lower weighted drive-cycle power losses, is

selected for a comparative analysis with respect to the original design.

The cross-sections and the flux line distributions under open circuit, and rated

load conditions of the two machines are shown in Fig. 6.25. Their main design

features are compared in Table 6.5, according to which the optimized design has also

a lower torque ripple and lower degree of PM demagnetization. As can be seen in Fig.

6.25, the saturation is more prevalent throughout the ferrous core of the optimized

design due to the stronger rotor field.

The efficiency map of the optimized design “D3” is provided in Fig. 6.26. The

efficiency map of the original design is also repeated in Fig. 6.26 for comparison. It

can be seen in Fig. 6.26 that the high efficiency contours of the optimized design

are directed towards the representative load points with the highest energy weights,

see Table 6.2, which are scattered along the high torque, and high speed vicinity.
183

100 95

80

Efficiency (%)
Torque (Nm) 90
60

40 85
20
80
2000 4000 6000 8000 10000 12000
Speed (r/min)

(a) D1

100 95

80

Efficiency (%)
Torque (Nm)

90
60

40 85
20
80
2000 4000 6000 8000 10000 12000
Speed (r/min)

(b) O

100 95
80
Efficiency (%)
Torque (Nm)

90
60

40 85
20
80
2000 4000 6000 8000 10000 12000
Speed (r/min)

(c) D2

Figure 6.21: Efficiency maps of the optimized spoke-type designs.


184

8
100

Copper losses (kW)


80 6
Torque (Nm)
60
4
40
2
20
0
2000 4000 6000 8000 10000 12000
Speed (r/min)
(a) D1

8
100

Copper losses (kW)


80 6
Torque (Nm)

60
4
40
2
20
0
2000 4000 6000 8000 10000 12000
Speed (r/min)
(b) O

8
100
Copper losses (kW)

80 6
Torque (Nm)

60
4
40
2
20
0
2000 4000 6000 8000 10000 12000
Speed (r/min)
(c) D2

Figure 6.22: Copper loss maps of the optimized spoke-type designs.


185

2
100

Core and PM losses (kW)


80 1.5
Torque (Nm)
60
1
40
0.5
20
0
2000 4000 6000 8000 10000 12000
Speed (r/min)
(a) D1

2
100

Core and PM losses (kW)


80 1.5
Torque (Nm)

60
1
40
0.5
20
0
2000 4000 6000 8000 10000 12000
Speed (r/min)
(b) O

2
100
Core and PM losses (kW)

80 1.5
Torque (Nm)

60
1
40
0.5
20
0
2000 4000 6000 8000 10000 12000
Speed (r/min)
(c) D2

Figure 6.23: Core loss maps of the optimized spoke-type designs.


186

60

40

Mass (kg)
20
O
D3
0
0 50 100 150
Weighted loss (pu)

Figure 6.24: Identifying a design with higher power density and drive-cycle efficiency
than the original design.

(a) Original, open circuit (b) “D3”, open circuit

(c) Original, rated load (d) “D3”, rated load

Figure 6.25: Flux lines distributions at 6000 r/min.


187

100 95

80

Efficiency (%)
Torque (Nm)
90
60

40 85
20
80
2000 4000 6000 8000 10000 12000
Speed (r/min)

(a) O

100 95

80

Efficiency (%)
Torque (Nm)

90
60

40 85
20
80
2000 4000 6000 8000 10000 12000
Speed (r/min)

(b) D3

Figure 6.26: Efficiency maps of the high power density spoke-type designs.

Furthermore, the optimized design has better torque production capability in the

extended speed range.

6.4 Summary

In the first case study, the large-scale CMODE-type design optimization approach

was successfully performed on the Toyota Prius Gen. 2 IPM traction motor, and

the results were verified through multi-physics performance analysis of the optimized

designs.
188

The optimization results confirmed the cost-effectiveness of the original motor

design. However, the large-scale multi-objective optimization was able to provide

alternative designs with higher drive-cycle efficiency, and with minimum additional

cost compared to the original design.

The second case study was performed on a high power density spoke-type

PM motor to further increase the power density of the original design. The

performance trade-offs associated with achieving high power density in such motors

were investigated. It was demonstrated that, in general, high power density is directly

correlated with higher losses, higher torque ripple, and larger PM demagnetization.

Furthermore, larger amounts of copper and PM are to be used in high power density

motors. The developed design optimization method was able to increase the power

density of the original design by 15%, and at the same time decrease the drive-cycle

power losses by 1%.


189

CHAPTER 7
ADDITIONAL IMPLICATIONS OF PM
MACHINES’ DESIGN OPTIMIZATION

In this chapter, three different aspects of design optimization of PM machines are

investigated.

In the first section, a numerical technique is developed for sensitivity analysis

of active material cost (AMC) in PM motors with distributed and fractional slot

concentrated windings. A comprehensive analysis is carried out to identify how the

optimal design rules and proportions of IPM motors with sintered NdFeB magnets

vary with respect to the changes in the commodity prices of permanent magnet

material, copper, and steel. The sensitivities of the correlations between the design

parameters and the AMC with respect to the commodity price ranges are investigated

based on response surface methodology (RSM) and large-scale design optimization

practice using differential evolution (DE) optimizer. An innovative application of

artificial neural network (ANN)-based design optimization is introduced. Multi-

objective minimization of cost and losses is pursued for an overall of 200,000 design

candidates in 30 different optimization instances subjected to different cost scenarios

according to a systematic design of experiments (DOE) procedure. An interesting

finding is that, despite common expectations, the average mass of steel in the

optimized designs is more sensitive to changes in the commodity prices than the

masses of copper and rotor PMs.


190

In the second section, a fast FE-based method for calculation of eddy current losses

in the stator windings of randomly wound electric machines will be presented. The

method is particularly suitable for implementation in large-scale design optimization

algorithms where a qualitative characterization of such losses at higher speeds is most

beneficial for identification of the design solutions which exhibit lowest overall losses

including the ac losses in the stator windings. Unlike the common practice of assuming

a constant slot fill (SF) factor for all the design variations, the maximum SF in the

developed method is determined based on the individual slot structure/dimensions

and strand wire specifications. Furthermore, in lieu of detailed modeling of the

conductor strands in the initial FE model, which significantly adds to the complexity

of the problem, an alternative rectangular coil modeling subject to a subsequent flux

mapping technique for determination of the impinging flux on each individual strand

is pursued. The research focus of the chapter is placed on the development of a

computationally efficient technique for the ac winding loss derivation applicable to

design optimization, where both the electromagnetic and thermal machine behaviors

are accounted for. The analysis is supplemented with an investigation of the influence

of the electrical loading on ac winding loss effects for a particular machine design, a

subject which has received less attention in the literature.

The last section of this chapter is dedicated to a revealing investigation of the

mechanical design of the rotor bridges in a typical single-layer V-type rotor PM layout,

similar to the one used throughout this dissertation. Particularly, the approximate

calculation of the centrifugal forces on the rotor bridges at high rotational speed

which is required to be estimated for adjustment of the width of the rotor bridges to,
191

300
Nd Cu Fe
21 250

Nd, Cu (USD/LB)

Fe ore (USD/t)
200
16
150
11
100
6
50

1 0
2008 2009 2010 2011
Year

Figure 7.1: Variation of the price of constructive commodities of Nd-based PM motors


for three consecutive years leading the surge of the Nd price in 2011 [4, 5].

on the one hand, withstand the maximum mechanical stress and, on the other hand,

minimize the leakage PM flux, which is magnetically shorted through these bridges,

will be discussed here. An analytical method to approximate the structural stress

due to centrifugal forces is implemented and its accuracy is compared to high fidelity

structural analysis using FEM.

7.1 RSM-DE-ANN Sensitivity Analysis of Mate-

rial Cost in PM Motors with Distributed and

Concentrated Windings

The price of active material from which PM motors are constructed, especially the

NdFeB- magnets, has experienced steep variations in the last few years. Figure 7.1

shows such variations over three consecutive years prior to the surge of the Nd price

in 2011 [4, 5, 182].

The minimization of such PM motor cost by eliminating the dependency on the


192

rare-earth PMs as in [183, 184], or by maximizing the effective utilization of costly

PM materials using optimization procedures as in [88, 185], has been the subject of

numerous studies. As for the latter case, a popular practice is to minimize the cost

associated to the active materials which are utilized in the machine construction.

Accordingly, a cost function is commonly defined as in (7.1.1) by aggregating the

masses of the active materials which are weighted by price coefficients corresponding

to their market values:

AM C = [ppm , pcu , pf e ] · [mpm , mcu , mf e ]T (7.1.1)

where [ppm , pcu , pf e ] and [mpm , mcu , mf e ] are the commodity price coefficient vector

and the commodity mass vector, respectively.

One of the early investigations which incorporated this cost model is the work

of Lovelace et al. [74, 186], according to which the cost factors of 5 USD per unit

weight (USD/w), 1 USD/w, and 11 USD/w were respectively considered for copper,

steel, and bonded Ferrite for IPM machines. Other investigators have adopted this

method of machine cost modeling with different price coefficients depending on the

market values and material specifications in their design optimization endeavors. In

their series of works, Zhang et al. assumed price coefficients of 8, 1, and 140 [85], or 8,

1, and 65 [88] for copper, lamination steel, and sintered NdFeB PMs, whereas Duan

and Ionel assumed price coefficients of 10, 1, and 100 [87], and Fatemi et al. assumed

price coefficients of 3, 1, and 24 [108–110] for the corresponding materials. Also in

[187], price coefficients of 6, and 140 were assumed for copper and PMs, respectively.

As stated above, based on the material specifications and market values, different
193

price coefficients have been incorporated in the definition of the cost model in Eq.

7.1.1. However, a systematic investigation of dependency of active material cost of PM

motors on the changes in the prices of such motors’ construction commodities, to the

best of these investigators’ knowledge, is not available in the literature. Furthermore,

although in all of aforementioned investigations various PM machine topologies were

optimized for cost minimization, the question still remains as to whether and how the

optimal design of such machines varies with respect to the assumed commodity price

coefficients.

In this regard, two sets of sensitivity analyses on two generic industrial IPM motor

topologies with distributed and concentrated winding configurations are conducted

in this chapter. In the first set, the change of the design parameters’ impacts on

the motor active material cost (AMC) due to variations in the price coefficients in a

typical cost function, similar to that expressed in Eq. (7.1.1), will be investigated.

In the second set, the sensitivities of the distribution of the design parameters in

the optimized candidate designs, which were obtained from large-scale optimization

processes, to the changes in the commodity price coefficients are presented.

7.1.1 FE-Based Machine Models

Two IPM machine configurations that were previously introduced in Section 3.2 are

chosen for this analysis. The vectors of the design parameters defined over the cross

sections of the two machines are given in Eq. (7.1.2), and Eq. (7.1.3) for the 48-slot

8-pole, 48S8P, and the 12-slot 10-pole, 12S10P, machines, respectively. These design

parameters and their bounds are identical to those previously described in Section
194

3.2.

X̄48S8P = (ksi , hg , kwt , kwtt , kdP M , kwP M , kwq , hP M , αP M , hy )T (7.1.2)

X̄12S10P = (ksi , hg , kwt , kdP M , kwP M , kwq , hP M , αP M , hy )T (7.1.3)

The two following performance metrics are evaluated using the FE models:

• Active material cost (AMC) according to Eq. (7.1.1).

• Power losses including dc copper losses, PM eddy current losses and core losses

at 100 ◦ C.

7.1.2 Effects of the Commodity Price on the Design

Correlations

The correlation between the AMC defined in Eq. (7.1.1) and the design vectors

given in Eq. (7.1.2) and Eq. (7.1.3) is bilateral; that is, on the one hand, for a

given set of commodity price coefficients, the design parameters determine the AMC.

On the other hand, the influences of the design parameters on the AMC are by

themselves a function of the commodity price coefficients. In this section, these

bilateral effects of design parameters and commodity price coefficients on the AMC

will be discussed. For this purpose, the information about the range of the changes

of the design parameters and the commodity price coefficients are required. The

range of the design parameters are designated in Table 3.1 whereby wide ranges are

assumed to fully explore the design space. As for the vector of the commodity price

coefficients, a range of [50, 4, 1] to [100, 8, 2] is assumed based on realistic variations


195

100

Regression coefficients
80
60
40
20
0
−20
ksi hg kwt kwtt kdPM kwPM kwq hPM α hy
PM

(a) 48S8P motor configuration

100
Regression coefficients

80
60
40
20
0
−20
ksi hg kwt kdPM kwPM kwq hPM α hy
PM

(b) 12S10P motor configuration

Figure 7.2: Influence of the design parameters on AMC in the two case-study motors.

of the utilized materials, i.e. sintered NdFeB magnets, electrical steel, and copper,

within the next few years.

In the first place, a sensitivity analysis is carried out to understand how the

changes in the design parameters influence the AMC for the mid-range of the

commodity price coefficients, i.e pP M = 75, pCu = 6, pF e = 1.5. The regression

coefficients of the second order response surface associated with AMC which are

given in Eq. (3.3.1) are calculated for the two 48S8P, and 12S10P configurations, and

are illustrated in Fig. 7.2(a) and (b), respectively.


196

Change of regression coefficients


20
Steel
15 Copper
PM
10

−5
ksi hg kwt kwtt kdPM kwPM kwq hPM α hy
PM

(a) 48S8P motor configuration


Change of regression coefficients

20
Steel
15 Copper
PM
10

−5
ksi hg kwt kdPM kwPM kwq hPM α hy
PM

(b) 12S10P motor configuration

Figure 7.3: Change of the influences of the design parameters on AMC based on the
value of the commodity price coefficients.

For both motor configurations, the split ratio, ksi , air-gap height, hg , tooth-

stem width, kwt , PM height, hP M , and yoke height, hy , show the strongest positive

correlation with AMC, indicating that an increase in these parameters would add to

AMC. It is interesting to note that despite the fact that the PM has the greatest

price coefficients, the design parameters directly associated with the mass of PM, i.e.

kwP M , hP M , and αP M constitute marginal contributors to AMC.

As mentioned previously, the results of the sensitivity analysis shown in Fig. 7.2
197

are derived for the mid-range of the commodity price coefficients. It is not understood

yet whether and how these correlations between the design parameters and AMC

would change depending on the variations of the commodity price coefficients. To

discern the influence of the commodity price variations on the correlations between

the design parameters and AMC, a second set of sensitivity analysis is in order. In this

case, the sensitivities of the correlations between the design parameters and AMC due

to the changes of the commodity price coefficients are investigated. For this purpose,

depending on the motor configuration, ten or nine second order response surfaces are

defined according to Eq. (3.3.1) for the correlations of each design parameter of the

48S8P and 12S10P motors, respectively. The regression coefficients for the correlation

of each design parameter are solved by a DOE study based on CCD method, which

is conceived this time for the coded commodity price coefficients. The results of this

analysis are illustrated in Fig. 7.3(a) and (b), for the 48S8P and 12S10P machines,

respectively.

Accordingly, it can be seen in Fig. 7.3 that the PM price, due to the greater

designated coefficients, followed by the steel price, owing to the larger masses of steel

used in the machine construction, show the strongest influences on the correlations

of the design parameters on AMC. Particularly, the design parameters which were

demonstrated to have the strongest correlations with the AMC for the medium range

of commodity price coefficients in Fig. 7.2 are affected most severely. For example,

according to Fig. 7.2, increasing the split ratio, ksi , translates into higher AMC.

Meanwhile, according to Fig. 7.3, the increase of the commodity price coefficients

associated with the PM or steel materials further intensifies the already positive
198

correlation between ksi and AMC. Similar trends exist for air-gap height, hg , tooth-

stem width, kwt , PM height, hP M , and yoke height, hy . That is, on the one hand,

increasing these parameters would increase the AMC according to Fig. 7.2, and on

the other hand, the increase of PM price, and steel price with the exception of hP M ,

would make these positive correlations stronger.

Similarly, by comparing the values of the regression coefficients in Fig. 7.3 with

those illustrated in Fig. 7.2, it is evident that the strongest changes of the most

influencing design parameters due to variations of the commodity prices are in the

same direction as the original effects of the design parameters. This fact suggests that

the optimal design parameters for AMC minimization should not differ substantially

if the price coefficients are changed.

It is also interesting to note that although copper has a higher price coefficient

in comparison to ferrous lamination materials, the variation of the copper price does

not affect the influence of the design parameters on the AMC in a meaningful way in

these case study design examples.

7.1.3 ANN-Based Design Optimization with Different Com-

modity Price Coefficients

The results of the sensitivity analysis in the previous section demonstrated that

the existing correlations between the design parameters and AMC are dependent

on the commodity price coefficients, particularly on the price of the PM material and

lamination steel. However, it is not clear yet to what extent the optimal ranges of

the design parameters for cost-effective design of PM machines change commensurate


199

with the variations of the active material prices.

To fully explore the degree of this dependency in a practical design problem,

where the interactions between various design parameters subject to a set of

objectives are involved, an evidence-based study of optimal design of cost-effective PM

machines through the large-scale optimization of such machines should be conducted.

Subsequently, a sensitivity analysis of the optimal design parameters with respect

to the commodity price coefficients, similar to what was performed in the previous

section, should be pursued. For this purpose, the optimal design parameters need

to be derived for a series of price coefficients which are prescribed by DOE. For

the three components of PM, copper, and steel, a second order response surface using

CCD method would require 15 distinct runs of large-scale optimization for each motor

configuration. To implement a design optimization with such a large extent, surrogate

modeling techniques for fast performance characterization of the design candidates

are desired. Here, artificial neural networks (ANN) are utilized [188] for this aim.

The ANN-based modeling has been previously investigated in the literature for

the design optimization of electromagnetic devices [189–191]. The application of

the ANN-based modeling is particularly suitable in the overall procedure for the

sensitivity analysis of the optimal design parameters with respect to commodity

price variations shown in Fig. 7.4. This is due to the fact that the same machine

configuration needs to be modeled in several runs of large-scale design optimizations,

here 15 runs for each example machine. The high fidelity FE-based machine models

which were developed in the previous section can be used during the first run of the

optimization. The resultant design members of this first optimization run can then be
200

Commodity Price Large-Scale Design Synthesizer


DOE
Price Range Coefficients
Loss, Cost Optimal Sensitivity
RSM
ANN-based DE-based Design Values Analysis
FE Machine Train ANN-based Model Design Optimizer
Model ANN Model Vector

Figure 7.4: Developed procedure for sensitivity analysis of the optimal design values.

utilized to train the artificial neural networks for fast modeling of each corresponding

machine configuration in the subsequent optimization runs.

Here, a two-layer feedforward network with one layer of 100 neurons with

hyperbolic tangent sigmoid transfer function, followed by an output layer of linear

neurons was created in MATLAB. The network was trained using the Levenberg-

Marquardt training method using the mean square error performance function [188].

The FE-based models developed in the previous section were used to train the

network for computation of the two performance metrics discussed in the previous

section. However, instead of calculating the AMC directly, the masses of the

constitutive components are obtained to be implemented in the AMC expression

in Eq. (7.1.1) with each set of commodity price coefficients, which are determined

by DOE, for each run of the optimization. The values of these performance metrics

calculated by the developed ANN-based method were compared to the values derived

from the FE-based model for a 1000 designs. This is in order to verify that the

estimation error of the ANN-based model is less than 1% for the two machine

configurations, as can be seen in Fig. 7.5.

The ANN-based models were incorporated into a large-scale design optimization

process with two main objectives:


201

1.5
mpm mcu mfe loss
1
0.5

Error (%)
0
−0.5
−1
−1.5
0 200 400 600 800 1000
Design number
(a) 48S8P motor configuration

1.5
mpm mcu mfe loss
1
0.5
Error (%)

0
−0.5
−1
−1.5
0 200 400 600 800 1000
Design number
(b) 12S10P motor configuration

Figure 7.5: Estimation error of the ANN-based models when compared to FE-based
machine models.

1. Minimization of AMC based on the set of commodity price coefficients generated

by DOE.

2. Minimization of power losses including core and copper losses which are

estimated using the ANN-based model.

The differential evolution (DE)-based optimizer [110] presented in Chapter 4,

was used as the stochastic search engine to identify the Pareto optimal designs.

Accordingly, 30 runs of optimization with various sets of price coefficients determined


202

by DOE were carried out. The optimization results for the mid-range price coefficients

expressed in terms of the objectives and color-coded with respect to the cost of the

utilized active materials are shown in Figs. 7.6 and 7.7 for the 48S8P and 12S10P

machines, respectively.

The correlations between the design objectives and the costs of the individual

constitutive active materials in Figs. 7.6 and 7.7 should be pointed out. The cost

component associated with the PM material has a strong positive correlation with

both high efficiency and AMC. That is, in order increase the efficiency, larger amounts

of PM is to be utilized, which in turn translates into higher overall AMC. Similar

trends exist between the steel mass and the optimization objectives, although the

correlations are not as strong as those attributed to the PM material. Meanwhile,

these correlations are reversed between the copper mass and the optimization

objectives. That is, as the utilized copper increases, the overall AMC decreases

whereas the total losses increase. It is also interesting to note that lesser amounts

of copper are required in the construction of the 12S10P machines when compared

to the 48S8P machine because of the shorter stack lengths and shorter end windings

of the 12S10P machine. However, the 12S10P machine suffers from slightly higher

power losses mainly due to the increased rotor core losses.

7.1.4 Sensitivity of Optimal Design Values to Commodity

Price Coefficients

In order to understand the dependency of the optimal design of PM machines

on the commodity price coefficients, a sensitivity analysis on the averages of the


203

700 300
600 250

PM cost (pu)
AMC (pu)

500
200
400
150
300
200 100
100 50
250 300 350 400 450 500 550
Power losses (W)

700 60
600

Copper cost (pu)


AMC (pu)

500 50
400
300 40
200
100 30
250 300 350 400 450 500 550
Power losses (W)

700 100
600 Steel cost (pu)
AMC (pu)

500 80
400
300 60
200
100 40
250 300 350 400 450 500 550
Power losses (W)

Figure 7.6: Optimization results of the 48S8P motor configuration for a typical set
of commodity price coefficients.
204

700 300
600 250

PM cost (pu)
AMC (pu)

500
200
400
150
300
200 100
100 50
250 300 350 400 450 500 550
Power losses (W)

700 60
600

Copper cost (pu)


AMC (pu)

500 50
400
300 40
200
100 30
250 300 350 400 450 500 550
Power losses (W)

700 100
600 Steel cost (pu)
AMC (pu)

500 80
400
300 60
200
100 40
250 300 350 400 450 500 550
Power losses (W)

Figure 7.7: Optimization results of the 12S10P motor configuration for a typical set
of commodity price coefficients.
205

design parameters in the optimized designs yielded by each round of optimization

is performed in this section. For this purpose, 100 of best design solutions with

minimized AMC and losses are identified from each set of optimization results.

For each set of the selected designs, the averages of the design parameters are

obtained together with the average masses of their active components. Subsequently,

a sensitivity analysis similar to what was described in Section 7.1.2 is conducted to

reveal how the changes of the commodity price coefficients influence the optimal range

of the design parameters. Figures 7.8(a) and (b) show the results of this sensitivity

analysis for the the 48S8P, and 12S10P motor configurations, respectively.

According to Fig. 7.8, for both machine configurations, the average values of the

optimal design parameters as well as the average masses of the PM material and

copper in the optimal designs are not sensitive to changes of the commodity price

coefficients. Meanwhile, the average masses of electric steel in the selected optimized

designs show the largest correlations with the price coefficients particularly since the

mass of the utilized steel in the machine construction is affected by all the design

parameters, and thus reflects the aggregate effects of the variations of the commodity

price coefficients. As can be seen in Fig. 7.8, an increase in steel price results in

lower average steel mass in the optimized designs for both motor configurations. The

impacts of changes of copper and magnet prices on the steel mass in the optimal

designs differ between the two motor configurations. Most notably, the steel mass is

expected to increase if the copper price coefficient in the 48S8P machine, or the PM

price coefficient in the 12S10P machine increases. Overall, a comparison between the

magnitudes of the regression coefficients in Fig. 7.8 with those provided in Figs. 7.2
206

0.5

Regression coefficients
0

−0.5
Steel
Copper
PM
−1
ksi hg kwt kwttkdPMkwPM kwq hPM α hy mFe mCumPM
PM
(a) 48S8P motor configuration

1
Regression coefficients

0.5

−0.5

−1
Steel
−1.5 Copper
PM
−2
PM αPM
k h k k k k h h m m m
si g wt dPM wPM wq y Fe Cu PM

(b) 12S10P motor configuration

Figure 7.8: The sensitivity of the optimal design parameters to variation of the
commodity price coefficients.

and 7.3 suggests that the optimal design values are relatively independent from the

assumed set of commodity price coefficients.

7.2 A Computationally Efficient Method for

Calculation of Strand Eddy Current Losses

The eddy current effects including the skin, strand-level, and bundle-level proximity

effects [192] potentially constitute a significant contributor to the overall copper


207

Figure 7.9: Slot leakage and fringing flux in a typical open-slot FSCW PM machine.

losses in the stator windings of high speed permanent magnet (PM) machines.

Even if preventive measures such as stranding and transposition are adopted, the

ac conductor losses in the stator windings of PM motors can still be significant for

high power density-high speed open-slot fractional-slot concentrated winding (FSCW)

machines due to the prevalence of slot leakage flux, and slot opening fringing flux,

e.g. see Fig. 7.9. Common techniques for the estimation of such losses are especially

prohibitive for randomly wound coil configurations and require a significant amount

of time to formulate and solve the ac electromagnetic problem at the conductor

strand level. The development of an FEA loss characterization method to provide

a basis of qualitative ac loss comparison between thousands of design candidates,

which is suitable for implementation in large-scale design optimization algorithms, is

imperative. In addition to the value of the ac loss, the distribution of the overall

copper losses in the stator winding is of interest particularly if coupled-thermal

electromagnetic design optimization is pursued [193, 194].

In a broad categorization, the popular methods for analysis of eddy current

losses in PM machines rely on analytical models as in [195–203], numerical finite


208

element/difference analysis as in [16, 204–206], or rely on combined analytical-FE-

experimental procedures as in [207–209]. The analytical models lack the desired

accuracy under magnetic core saturation and are not applicable to complex geometry

without compromising further the accuracy. The numerical models are not suitable

for integration into large-scale design optimization processes due to time consuming

computations. The combined procedures require extensive a priori experimentations

and are best suited for accurate loss analysis between different motor and winding

configurations as opposed to application for large-scale design optimization of one

particular configuration.

In this section, a finite element (FE)-based modeling technique is presented for

estimation of the strand eddy current losses in the stator windings of electric machines,

with an emphasis on sinusoidally excited PM synchronous machines. This is in order

to include the portion of the ac losses which stems from the presence of slot leakage and

fringing fluxes in the performance characterization of such machines. The developed

hybrid analytical-numerical loss calculation method is rendered computationally

efficient through adopting several measures such as alternative coil modeling which

reduces the computation time required for solving the FE model, exploiting the

existing electric symmetry in addition to the magnetic periodicity of PM machines

with sinusoidal current excitation, and implementing fast analytical techniques for

mapping the flux within the slot area, estimating the fill factor and strand locations,

and finally characterization of the eddy current losses based on the value of the flux

density impinging on each stator winding conductor. The presented loss calculation

method provides a reasonable compromise between computational time and accuracy,


209

which makes it suitable for application in large-scale design optimization of PM

machines in the initial stages of the design.

Using the developed method, strand eddy current losses under various loading

conditions are computed and the existing trends between the ratio of ac to dc losses,

Pac /Pdc , with respect to the loading level are studied. Through this analysis, it

is demonstrated that the traditional figure of merit for comparison of ac losses in

electric machines, which is established by the ratio of ac to dc resistance, Rac /Rdc ,

is by definition incapable of modeling the effects of the loading level on the ratio of

Pac /Pdc .

7.2.1 Strand Eddy Current Loss Characterization

Analytical models are reported for (a) 1-D single-slot models as in [197], (b) 2-D

single-slot models as in [198], or (c) 2-D machine models as in [199]. These methods

provide an insight into the nature of eddy current losses but do not accurately account

for the non-linearity of the magnetic core, and are difficult to apply to complex

machine geometries.

Numerical models require a significantly large number of elements in the stator

slots and are therefore time-consuming. In some studies with detailed coil models

[16, 205], an even distribution of current is assumed in the conductors so that a

time-stepping magnetostatic solution can be performed. Subsequently, a detailed

distribution of the radial and tangential components of the flux density, BR,T , in each

stator slot is obtained by establishing a fine grid over each slot pitch of the stator from

one tooth axis to the next. These values are used in a numerical harmonic analysis
210

expressed by:


X 
BR,T (t) = |B2k−1,(R,T ) | sin (2k − 1)ωt − φ2k−1,(R,T ) (7.2.1)
k=1

Accordingly, the eddy current loss, Pe , in watts per strand per depth of axial

length, for a rectangular copper strand of width a, and height b subject to a uniform

time varying flux density of Eq. (7.2.1) can be obtained using [199]:


X (2k − 1)2 ω 2
|B2k−1,R |2 a2 + |B2k−1,T |2 b2 η2k−1

Pe = ab (7.2.2)
k=1
24ρ

where the skin effect coefficient η2k−1 is given by [210]:


s
3 sinh(2α2k−1 ) − sin(2α2k−1 ) a µr (2k − 1)f
η2k−1 = 2
, α2k−1 = (7.2.3)
4α2k−1 cosh(2α2k−1 ) − cos(2α2k−1 ) 6320 ρ

In the case of round conductors of diameter d, and resistivity ρ, Eq. (7.2.1)

can be used for calculating the magnitude of the impinging flux |B2k−1 | =
p
|B2k−1,R |2 + |B2k−1,T |2 , with the eddy current loss per depth of axial length given

by [192]:

4
X (2k − 1)ω 2 |B2k−1 |2
Pe = πd (7.2.4)
k=1
128ρ

The ratio of ac to dc resistance, krac , of round conductors can be modified

according to Eq. (7.2.5) in order to also include the skin effect [192]:
!
d
r
d I0 ( 2δ (1 + j)) 2ρ
krac = Re (1 + j) d
,δ = (7.2.5)
8δ I1 ( 2δ (1 + j)) ωµ

where I0 and I1 are Bessel functions of zero and first orders, respectively, and δ is the

skin depth. The solution of rac for round conductors is documented through charts

and graphs in [192], and can be readily found for a given conductor diameter and

excitation frequency.
211

7.2.2 FE-Based Eddy Current Loss Estimation for Randomly

Wound Stator Windings

For a given machine configuration, the distribution of the leakage/fringing flux within

any slot is dependent on:

• various dimensions of the cross section,

• the loading level,

• the location of the conductors within the stator slots,

• the temperatures of various components, and

• the frequency of operation.

Only a thermally coupled-time-harmonic finite element (FE) model with detailed

knowledge of conductor locations can account for all the aforementioned parameters.

The formulation and solution of such an electromagnetic (EM) problem is an

extensively and computationally demanding process, not suitable for early design

stage optimization purposes.

In this section, an alternative method will be presented. The steps of this

computationally efficient loss calculation method which can be integrated into a large-

scale design optimization process are described in the following section.

7.2.2.1 Modeling of the Coils

Detailed modeling of the coils in the slots, such as the one shown in Fig. 7.10(a), adds

to the complexity of the FE model, and thus increases the computation time to reach
212

(a) Detailed model (b) Commonly used model (c) Alternative model

Figure 7.10: Alternative coil models for strand eddy current loss analysis.

a solution. As opposed to the crude coil model commonly used for EM-FE analysis,

shown in Fig. 7.10(b), here an alternative representation is developed. According to

Fig. 7.10(c), the winding is divided into a number of rectangular areas over its radial

and tangential dimensions.

The heights and the widths of the sections can be all equal or skewed to provide

more details at the slot opening. In Fig. 7.10(c), 32 sections with equal heights

and equal widths are defined. The number of sections should be selected so that a

reasonable distribution of B samples are obtained within the slot. The radial and

tangential components of the flux densities at the middle of each rectangular section

can then be extracted. The sampled B profiles are subsequently used to map the flux

density at any point in the slot.

7.2.2.2 Extraction of B Field Profiles Inside the Slot

The value of B at the middle of each section can be obtained by any time-stepping

magnetostatic FE analysis including the computationally efficient FEA (CE-FEA)

method introduced in [81, 119], which exploits the electric symmetry in the stator
213

windings of sinewave operated/energized PM motors. As described in Section 2.1.1,

using CE-FEA, the profile of the flux density waveforms over the full electrical cycle

can be reconstructed by performing FEA over a window of 60 elec. deg.

Here, the CE-FEA method is used to extract the radial and tangential components

of the sampled B profiles for the coil pieces shown in Fig. 7.10(c) for a typical machine

under full-load motoring operation with counterclockwise rotation. The profiles for

selected sections are shown in Figs. 7.11(a) through (d). It is interesting to note

that the major component of the slot leakage flux is tangential. Furthermore, the

decreasing trend of this slot leakage flux from top to bottom of the slot, and from left

to right for the motoring operation, should be noted.

7.2.2.3 Determination of the Slot Fill Factor and the Associated

Conductor Positions

The slot fill factor, sf , is defined as follows:

ACu nC πd2
sf = = (7.2.6)
Aslot 4Aslot

where ACu is the copper area within the slot area, Aslot , and nC is the number of

conductors. In a large-scale design optimization problem, the slot dimensions and area

vary between the design candidates. Accordingly, the maximum slot fill factor, sf,max ,

and the location of the conductors vary between the design candidates and should

be calculated for each individual design. Here, the sf,max is needed for the FEA to

accurately account for the available Ampere Turn flowing through the alternate coil

model described in the previous section. Furthermore, the conductor positions inside

the slot are required for determining the impinging field on each strand in a post-FEA
214

0.1
1
2
0.05

Radial B(T)
3
4
0
5
6
−0.05
7
8
−0.1
0 60 120 180 240 300 360
Rotor Position (deg. el.)
(a) Radial component of B moving into the slot

0.1
1
0.05 19
Radial B(T)

17
0 25

−0.05

−0.1
0 60 120 180 240 300 360
Rotor Position (deg. el.)
(b) Radial component of B moving along the slot opening

0.3
1
0.2 2
Tangential B(T)

0.1 3
4
0
5
−0.1 6
−0.2 7
8
0 60 120 180 240 300 360
Rotor Position (deg. el.)
(c) Tangential component of B moving into the slot

0.3
0.2 1
Tangential B(T)

19
0.1 17
0 25
−0.1
−0.2

0 60 120 180 240 300 360


Rotor Position (deg. el.)
(d) Tangential component of B moving along the slot opening

Figure 7.11: Radial and tangential components of the field sections in Fig. 7.10(c)
for a typical motor.
30

215
25

20

15

10

−5

Figure 7.12: Determination


−10
of sf,max and conductor positions by moving the slot
geometry over a grid of conductors.
−15
60 65 70 75 80 85 90 95 100 105 110

process using the B samples from FEA. This mapping process will be described in a

later section.

The method that was implemented here for the calculation of sf,max and the

associated positions of the conductors within the slot, which yields such sf,max ,

relies on an optimization approach that is based on random perturbation of the slot

geometry [117]. As can be seen in Fig. 7.12, a given slot geometry is randomly moved

with respect to a grid of tightly packed circular conductors. The fill factor is then

compared for various slot perturbations and strand arrangements to determine the

sf,max and the associated locations of the strands.

This method is used for the determination of sf,max in the example slot geometries

shown in Figs. 7.13(a) through (c). As can be seen in this illustration, for a given

conductor diameter, the achievable sf,max diminishes as the slot area decreases, which

is successfully predicted by the method implemented here.


35

35 35
216
30

30 30

25
25 25

20
20 20

15
15 15

10
10 10

5 5
5

0 0 0

−5
(a) sf,max = 0.47 −5
(b) sf,max = 0.51 −5
(c) sf,max = 0.53
−10 −10 −10
Figure 7.13: Slot fill factor and strand positions for example slot geometries as the
−15
60 net slot area increases.
65 70 75 80 85 90
−15
95 60 10065 10570 110 75 80 85 90 95
−15
60
100
65
105
70
110
75 80 85 90 95 100 105 110

7.2.2.4 Mapping the Flux on Each Individual Strand

When the B profiles over the full fundamental cycle are obtained for each rectangular

coil section, the radial and tangential components are separately used in a time

harmonic analysis according to (7.2.1). Subsequently, for each harmonic, a Delaunay

triangulation method is implemented in MATLAB for interpolating the samples

scattered over the slot at the points where the center of each strand conductor is

located. This process is illustrated in Fig. 7.14 for the reconstruction surfaces of the

first harmonic of the B field throughout the slot area.

The order of harmonics that should be included is design dependent. When the

field values throughout the slot area are determined, using the prior information of

the strand positions, the impinging field on each strand can be mapped. In Fig. 7.15,

the mapped values of the impinging B over each strand are shown for the fundamental

and the third field harmonics. It can be seen once again that the magnitude of the

slot leakage flux density monotonically increases for the strands that are closer to the

air gap. The same trend exists for strands that are located towards the leading end

of the rotor pole under motoring operation for CCW direction of rotation.
217

Figure 7.14: Reconstruction of the field harmonics from the sample points using
Delaunay triangulation method.

7.2.2.5 Estimation of Value and Distribution of Eddy Current Losses

Upon derivation of the impinging |B| on each strand, depending on the conductor

shape, the loss models given in Eq. (7.2.2) through Eq. (7.2.5) can be used for

estimation of strand eddy current losses in the conductors at the strand level. The

resultant loss values using such an analysis on a typical slot is shown in Figs. 7.16(a)

through (f) for a wide range of loading levels.

7.2.3 Case-Study Analysis

The method developed here was used for the calculation of eddy current losses in a

12-slot 10-pole IPM machine with V-type magnet layouts as shown in Fig. 7.17(a).

The all-teeth-wound stator winding consists of series coils each composed of 53 turns

of AWG 12.5 wires, thus reducing to negligible levels the losses associated with
218

(a) Typical slot and strand numbers

0.25

0.2
Impinging B(T)

0.15

0.1

0.05

0
20 40 60 80 100
Strand identifier
(b) The fundamental component of the impinging B

0.03
Impinging B(T)

0.02

0.01

0
20 40 60 80 100
Strand identifier
(c) The third harmonic component of the impinging B

Figure 7.15: Mapped flux on each individual strand.


219

(a) No-load (b) 25% load (c) 50% load

(d) 75% load (e) 100% load (f) 125% load

Figure 7.16: Distribution of strand eddy current losses under various loading levels.

circulating currents, which are essentially of a three-dimensional nature and cannot

be accounted for by two-dimensional models.

The stator winding losses including the strand eddy current losses are calculated

by the developed method and the results are compared with those obtained from a

time-stepping FEA with detailed coil modeling as shown in Fig. 7.17(b).

The strand eddy current losses are calculated over a wide range of motor loading

conditions under MTPA control for three different speeds at a winding temperature

of 100◦ C. The results obtained from the developed method and those from the time

harmonic FEA with detailed coil modeling are compared in Figs. 7.18(a)-(c). The

required computation time is less than 80 seconds using the proposed method as

opposed to 3370 seconds using the detailed FEA.

The estimation error of the proposed method when compared to the detailed
220

(a) CE-FEA with flux mapping (FM)

(b) FEA with detailed coil modeling

Figure 7.17: Case-study investigation of strand eddy current losses.


221

10
TS-FEA with Detailed Coil Modeling

EDDY LOSSES (W)


STATOR WINDING
8 CE-FEA with Flux Mapping

0
0 25 50 75 100 125
LOAD LEVEL (%)
(a) 150 Hz

120
TS-FEA with Detailed Coil Modeling
EDDY LOSSES (W)
STATOR WINDING

100
CE-FEA with Flux Mapping
80
60
40
20
0
0 25 50 75 100 125
LOAD LEVEL (%)
(b) 500 Hz

600
TS-FEA with Detailed Coil Modeling
EDDY LOSSES (W)
STATOR WINDING

500 CE-FEA with Flux Mapping


400
300
200
100
0
0 25 50 75 100 125
LOAD LEVEL (%)
(c) 1200 Hz

Figure 7.18: Comparison of the accuracy of the loss calculation method over a wide
range of frequencies and loading conditions.
222

15
150 Hz. 500 Hz. 1200 Hz.
10

ERROR (%)
5
0
-5
-10
-15
0 25 50 75 100 125
LOAD LEVEL (%)

Figure 7.19: Estimation error of the computationally efficient method of calculation


of strand eddy current losses compared to the full-fledged time harmonic analysis
with detailed coil modeling.

TS-FE model is shown in Fig. 7.19 for several speeds and over a wide range of

loading. The error is within a reasonable range given the computational efficiency of

the proposed method.

The variation of the ac to dc loss ratio, Pac /Pdc , due to the armature reaction

under different loading levels is shown in Fig. 7.21. As can be seen in Figs. 7.21(a)

through (c), strand eddy current losses constitute a larger contribution to the overall

losses, Pac = Pdc + Pe , under light load levels. The rate of increase of eddy current

losses with respect to loading, which is mainly due to the elevated saturation level of

the ferrous core and therefore increased leakage and fringing of flux into the slot area,

is less than the rate of increase of dc copper losses Pdc , which is directly proportional

to the current squared. This is especially true at lower frequencies as can be seen in

Fig. 7.20. However, the eddy current losses are constantly present even at no-load

conditions due to the presence of the time-varying field in the slots.

The ratio of Rac /Rdc , which is commonly used in the literature, is by definition

not exposed to such large variations, and thus does not reflect them. Therefore, if the
223

600
Eddy losses 150 Hz.
500

OHMIC LOSS (W)


Eddy losses 500 Hz.
400 Eddy losses 1200 Hz.
300 dc loss
200
100
0
0 25 50 75 100 125
LOAD LEVEL (%)

Figure 7.20: Variation of dc ohmic losses and strand eddy current losses with respect
to loading level.

ratio of Rac /Rdc is to be used as a figure of merit for comparison of ac losses between

different design solutions, it should be derived and formulated under various loading

conditions.

7.3 Estimation of Tangential Mechanical Stresses

on the Rotor Bridges

In this section, the analytical method introduced in [211] is used to develop relevant

formulations for the calculation of mechanical stresses in the bridges of a single-

layer V-type IPM rotor layout. The results are compared to those obtained from

the mechanical finite element analysis of the rotor structure. The comparison is

performed through a sensitivity analysis of the stresses on the bridges of the case-study

rotor using the two analytical and structural FE techniques for stress estimation.

Although the emphasis is placed on the single layer V-type rotor layout which

has been used throughout this dissertation, the methodology is applicable to many

rotor configurations commonly seen in axially laminated rotary IPM machines. The
224

1.25
AC TO DC LOSS RATIO 150 Hz.
1.2

1.15
y = 1.5837x-0.092
1.1

1.05

1
25 50 75 100 125
LOAD LEVEL (%)
(a) 150 Hz

3.5
AC TO DC LOSS RATIO

500 Hz.
3

2.5

2 y = 15.798x-0.527

1.5

1
25 50 75 100 125
LOAD LEVEL (%)
(b) 500 Hz

13
AC TO DC LOSS RATIO

1200 Hz.
11
9
7
y = 177.73x-0.878
5
3
1
25 50 75 100 125
LOAD LEVEL (%)
(c) 1200 Hz

Figure 7.21: Ratio of ac to dc losses over a wide range of loading conditions.


225

purpose of this development is to approximate the required adjustments of the widths

of the rotor bridges of various design candidates with different rotor parameters in a

large-scale design optimization process, and to account for the influence of the PM

leakage flux shorted through the rotor bridges on the electromagnetic performance

of the design solutions. The purpose of this analysis is not to omit the post-design

optimization investigation of the mechanical stresses followed by adopting necessary

measures for reducing the magnitude and/or the concentration of mechanical stresses

throughout the rotor geometry.

7.3.1 Adopted Analytical Stress Estimation Method

Numerous investigators have reported analytical procedures for calculation of

mechanical stresses in the rotor of the surface mounted permanent magnet machines.

These procedures are used to ensure adequate enclosure contact pressure is maintained

at high rotational speeds [176] in order to keep the magnets from flying off the rotor

surface. These analytical developments are based on stress formulations in rotating

concentric cylinders with different boundary equations with respect to the type of

the contact surfaces [212, 213]. Such analytical derivations do not exist for complex

geometries of IPM machines [159] without further sacrifice on accuracy of calculations

of the maximum stress [214]. Furthermore, analytical techniques are unable to yield

the distribution of the stress over the machine geometry. Still, a few attempts have

been made to estimate the maximum stress values for IPM machines which suit the

initial stages of the design process.

According to the method developed in [211], the tangential stress, σt , in the rotor
226

requiv,o
Aequiv
requiv,i
AFe,o

Apm1 Apm2

Original layout Equivalent ring

Figure 7.22: Modeling of the original layout on the left with the equivalent ring on
the right.

bridges of an IPM machine can be estimated using an equivalent ring arrangement as

sown in Fig. 7.22. In this approach, a hypothetical ring with is assumed. The height

of this equivalent ring is equal to the narrowest height of the rotor iron bridges.

Furthermore, it is characterized by an increased mass density, Dequiv , calculated

according to Eq. (7.3.1) as follows:

DF e · AF e,o + Dpm · Apm


Dequiv = (7.3.1)
Aequiv

where DF e is the mass density of rotor steel laminations, Dpm is the mass density of

PMs, AF e,o is the area of the iron under each pole piece, and Apm is the area of the

PMs per pole.

Subsequently, it can be assumed that the tangential stress inside the equivalent

rotating ring, σt,equiv , expressed in Eq. (7.3.2), is an indication of the tangential stress

present in the original layout. Eq. (7.3.2) is obtained based on formulations of the
227

hoop stress in rotating cylinders which are well-developed in the literature [176, 213].

requiv,o + requiv,i 2 2
σt,equiv = · ωmax · Dequiv (7.3.2)
2

where requiv,o and requiv,i are the outer and inner radii of the equivalent ring,

respectively.

To account for the stress concentrations due to the irregular magnet slot shape,

the calculated stress can be multiplied by a form factor, e.g. a factor of 2 as reported

in [211] provided that the sharp corners are rounded. The thickness of the rotor

bridges can be increased to reduce the stress on the rotor bridges to values less than

the yield strength of the laminations [214].

7.3.2 Evaluation of the Accuracy of the Analytical Stress

Estimation Method

In a large-scale design optimization, the methods utilized for modeling various

performance aspects of the design candidates should be capable of accurately relating

the design parameters to the desired performance metrics over the entire design space.

As mentioned in the previous discussion, the analytical methods for calculation of

the mechanical stresses in the rotor of IPM machines, do not reveal the concentration

and the distribution of the stresses throughout the rotor geometry. Furthermore,

many approximations are involved in derivation of such formulations. Therefore, it

is imperative to systematically compare these methods with high fidelity FE models

before employing them in the design practice.

Here, to understand the effectiveness of the method described in the previous

section in identifying the existing correlations between various design parameters and
228

maximum stress in the rotor bridges of a V-type PM machine, two rounds of sensitivity

analysis are performed. One uses the high fidelity structural FE model, and the other

one uses the aforementioned analytical method as means for stress calculations.

In preparing the FE model, the sharp edges in the rotor geometry are slightly

rounded, as shown in Fig. 7.23 for the two extremes of the rotor design parameters,

to avoid large stress concentration values in the ribs. The radius of filleting is

deliberately chosen to be small to minimize the effects on the machine electromagnetic

performance.

The sensitivity analysis methodology described in Section 7.1.2 is used with the

same range of design parameters associated with the rotor geometry indicated in

Table 2.1, i.e. kdpm , kwpm , kwq , hpm , αpm , in addition to rro which is defined by

ksi and hg parameters. A total of 45 distinct designs identified by the Design of

Experiments procedure using Central Composite Design methodology were analyzed

using both analytical and structural FEA. The stress magnitudes and distributions

of these designs obtained from structural FEA are provided in Appendix II. The

stress found from the analytical method, and the maximum stress from the structural

FEA are used for formulating two second order response surfaces expressed in Eq.

(3.3.1). The normalized regression coefficients obtained from the two methods of

stress calculation are shown in Fig. 7.24.

A comparison between the regression coefficients in Fig. 7.24 reveals that the

analytical method is successful in modeling the effects of the design parameters on

the bridge mechanical stresses except for two cases, namely, the PM depth ratio, kdpm ,

and the web width ratio, kwq . According to the analytical formulations, increasing
229

(a) Minimum values of the design parameters

(b) Maximum values of the design parameters

Figure 7.23: Preparing the rotor geometry for structural FE analysis.


230

Regression coefficients
0.8
0.6
0.4
0.2
0

−0.2
hpm kd kw kw αpm rro
pm pm q

(a) Minimum values of the design parameters


Regression coefficients

0.8
0.6
0.4
0.2
0

−0.2
hpm kd kw kw αpm rro
pm pm q

(b) Maximum values of the design parameters

Figure 7.24: Comparison of the accuracy of the analytical stress estimation method
through a sensitivity analysis.

the PM depth ratio when other design parameters are constant will increase the mass

under the pole piece, and thus elevates the stresses on the rotor bridges. This would

have held true had it not been for the existing features in the rotor geometry which

are not accounted for in the analytical model. Particularly, as the magnet burial

depth increases, the magnets are better retained from the centrifugal forces because

of the tilted inner edges at the two ends of the rotor slots marked in Fig. 7.25.

The analytical approach is also unable to effectively model the influence of kwq on
231

Figure 7.25: Change of kdpm ratio from lower to upper bound.

Figure 7.26: Change of kwq ratio from lower to upper bound.

the bridge stresses according to Fig. 7.24. This is due to the fact that the changes in

the mass of the pole piece due to variations of kwq are marginal as shown in Fig. 7.26.

Overall, the analytical approach provides an acceptable indication of stress which

suits the large-scale design optimization.


232

7.4 Summary

In this chapter, three additional aspects of design optimization of PM machines were

discussed:

In the first part, the sensitivity of the optimal design rules of IPM machines

with sintered NdFeB magnets to the variations of the commodity price coefficients

within a practical range of change was investigated in this section. Two sets of

rigorous sensitivity analysis were performed for this purpose. First, a comprehensive

sensitivity analysis was conducted on the impact of the commodity price variations

on the relationships between design parameters and the AMC. The results of this

sensitivity analysis demonstrated that the strongest changes of the most influential

design parameters due to variations of the commodity prices were in the same

direction as the original effects of these parameters.

In a second round of analysis, a comprehensive procedure based on a combined

RSM-DE-ANN technique was developed for large-scale multi-objective ANN-based

design optimization of over 200,000 design candidates to investigate the change in the

optimal design values due to the commodity price variations. The results indicated

that the average optimal design parameters are not prone to significant changes due

to the variations of the commodity price coefficients within the assumed ranges.

Similar conclusions would be expected for PM machines with Ferrite or bonded NdFeB

magnets, since the use of less expensive magnet materials translates into a lower price

coefficient, and thus reduces the influence of the PM price variations on the overall

active material cost.


233

In the second part of this chapter, a method was developed for the calculation

of strand eddy current losses in the stator windings of electric machines that (a) is

finite-element based to take into account the complex geometry of the machine and

the effects of saturation, (b) is computationally efficient and suitable for integration

into large-scale design optimization algorithms, (c) is applicable to any variety of

machines with different combinations of stator slots and rotor pole structures, (d)

estimates the maximum SF factor for each design candidate based on winding specs

and slot geometry, (e) estimates the value of eddy current losses due to slot leakage

and fringing flux effects under any loading conditions, i.e. various torque and speed

operating points, and (f) estimates the distribution of copper losses including the

eddy current losses in the slots for rigorous thermal analysis of the stator windings.

The developed loss calculation method was implemented on a FSCW 12-slot 10-

pole IPM machine with relatively large slot openings. The results over a wide range

of loading conditions and operating frequencies were in good agreement with those

obtained from a time harmonic FEA with detailed coil modeling. Meanwhile the

required computation time was significantly reduced using the presented method.

The distribution of the losses in this case study machine can be used for a subsequent

thermal performance analysis due to the importance of including strand eddy losses as

a major loss component in high speed machines, even if the stator winding conductors

are stranded and transposed.

Using the developed loss calculation method, it was also shown that the variations

of Pac /Pdc loss ratio with reference to the machine loading levels are not reflected in

the common figure of merit represented by Rac /Rdc resistance ratio. Thus if the
234

Rac /Rdc ratio is to be used, additional treatment will be required to include the

loading effects.

In the third and last part of this chapter, an analytical method was implemented

for the estimation of mechanical stresses in the rotor bridges of V-type IPM machines.

Through conducting rigorous sensitivity analyses using analytical and numerical

modeling tools, the capability of the utilized methodology in accurately describing

the relationships between the design parameters and the centrifugal stresses on the

rotor bridges were quantified. In the case study analysis, as expected, the analytical

modeling approach which was derived based on stress formulations in rotating

cylinders was not able to account for the effects of two of the design parameters,

namely the burial depth of PMs and the width of the q-axis webs, on the stress

concentrations in the rotor structure. However, the effects of the remaining four

parameters were predicted with reasonable accuracy using the analytical method,

thus rendering this approach one of the best available options for implementation in

large-scale design optimization procedures.


235

CHAPTER 8
CONCLUSIONS

8.1 Summary and Conclusion

In this dissertation, the following aspects of large-scale design optimization of

permanent magnet synchronous machines were presented:

In Chapter 1, the significance of research on high performance motor drive systems

and the challenges and opportunities for efficiency improvement in such systems were

discussed. In addition, the existing literature on numerical analysis and model-

based optimization of permanent magnet synchronous machines was reviewed. It

was pointed out that there is need for a more comprehensive design optimization

approach which is inclusive of the entire range of possible operating conditions. The

organization of this dissertation vis-a-vis the various investigated topics was also

outlined in this chapter.

In Chapter 2, a recently introduced model-based methodology for design synthesis

of synchronous PM machines was presented in detail. This approach, which

features computationally efficient-finite element analysis for the characterization of

performance of the design candidates, and utilization of the differential evolution

search algorithm for finding the globally optimal designs, was carried out for large-

scale design optimization of the Toyota Prius Gen 2 48-slot 8-pole V-type IPM

machine at its peak (nominal) operating condition. The optimization was performed
236

over 4 800 design candidates for realizing two objectives of minimization of active

material cost and power losses, in addition to satisfying two performance constraints

on maximum torque ripple and maximum permissible degree of PM demagnetization.

It was shown that more efficient and cost effective designs with superior rated

performance indices at the nominal (peak) operating point could be found from the

Pareto front of the optimization results. However, when the efficiency throughout the

entire range of operation is considered, the original Prius design outperformed the

selected Pareto optimal designs.

Demonstrating the necessity of including the entire range of operation in the

optimization process is one key contribution of this research which was described in

Chapter 3. Through a rigorous sensitivity analysis, it was shown that the correlation

indices between the design parameters and the performance metrics vary with respect

to the machine loading levels. In some instances, e.g. for the core and copper

losses, torque ripple, and degree of PM demagnetization, these correlation indices

undergo significant variations, or even change direction. These results put into

perspective the considerable amount of research in the literature, which based on

performance characterization at one operating point, prescribe various sizing/scaling

ratios for achieving or alleviating desired or undesired performance metrics. In a

second set of analyses in this chapter, six parallel design optimization runs consisting

of an overall of 40 000 design candidates were carried out at three different levels

of current densities, which are typically found in naturally cooled, fan-cooled, and

liquid-cooled machines. In addition, the analysis was conducted on two different

winding configurations, namely distributed and fractional slot concentrated windings.


237

The statistical distributions of the design parameters in the 500 optimal designs of

each optimization run were subsequently investigated. The results indicated that the

optimal design of PM machines vary with respect to the ampere loading level. This

highlight the challenges in the design of electric motors with wide operating ranges,

such as those used in traction applications, where an optimal design is to maintain

high performance under various complex patterns of loading conditions.

Taking into account the entire range of complex operation patterns in the

optimization fitness function contributes to the computational complexity of the

modeling process, especially when FE-based models are utilized. Accordingly,

a second key contribution of this dissertation is to mitigate the computational

complexity of the design optimization process. For this purpose, in Chapter 4, a

fast search algorithm was developed in this work for design optimization of electric

machines. Namely, this new combined multiobjective optimization with differential

evolution (CMODE-type) algorithm is best suited for implementation on multi-core

workstation computers owing to its distinctive steady-state evolution model, which

requires a lesser number of simultaneous function evaluations when compared to the

standard DE or GA. This CMODE-type algorithm was implemented here and was

thoroughly examined, and its superiorities over the standard DE algorithm in terms of

convergence rate, constraint handling, and quality of the final generated Pareto fronts

in a multiobjective design problem were confirmed and quantified. For this purpose,

12 independent runs of optimization, each consisting of 3 200 function evaluations,

were conducted on different machine topologies with various loading levels and fitness

functions, using either CMODE or DE as the search algorithm. Both counterpart


238

algorithms were able to identify the same global optima for all the optimization cases.

However, in all of the examined cases, CMODE’s convergence to the Pareto optimal

vicinity was faster, at least twice as fast as DE. Furthermore, CMODE’s constraint

handling was more effective in the sense that a larger number of design candidates

passed the designated performance constraints when compared to results from the

design space produced by DE. Thus, a salient finding in this dissertation is that these

characteristics of the CMODE algorithm render it the preferred search algorithm

for implementation on desktop workstation computers when a limited number of

processing cores and software licenses are available.

Another main contribution of this dissertation is the algorithm for large-scale

design optimization of PM machines over a target operating cycle which was presented

in Chapter 5. The resulting optimization process consists of the identification

of the motor torque and speed profiles, computationally efficient modeling of the

load energy distribution function, FEA-based performance evaluation at the cyclic

representative load points residing in the constant torque or extended speed regions,

and finding the optimal design solutions using the CMODE-type stochastic search

algorithm. The original CE-FEA algorithm was upgraded to include any load

operating point residing anywhere in the constant torque or extended speed regions

of the torque-speed plane. Furthermore, a new k-means clustering algorithm was

implemented for efficient modeling of the motor energy distribution function. Proper

designation of the objectives and constraints of the optimization fitness function

was also conceived in this chapter. It was demonstrated that the equality of the

characteristic and rated currents should not be considered as an objective of the design
239

optimization. Alternatively, a number of performance criteria should be pursued,

including improving upon the energy efficiency, while simultaneously checking the

torque production capability at the required load points.

The optimization algorithm which was developed in Chapter 5 is applicable to

the large family of sine-wave driven radial flux synchronous PM and synchronous

reluctance machines over any conceivable operating cycle. Another contribution is

that in Chapter 6, this approach was successfully applied to two case study traction

propulsion motors. In the first case study, the Toyota Prius Gen 2 motor was

optimized for reduced active material cost and increased drive-cycle efficiency over

a combination of common US driving schedules. By investigating roughly 10 000

design candidates over seven representative load points, the cost effectiveness of the

original design was confirmed and alternative designs with better energy efficiency

characteristics were identified. The viability of the final counterpart designs were

examined in a multiphysis performance analysis, including transient thermal modeling

over the rigorous US06 driving cycle, and mechanical FEA of the rotor structure at the

maximum rotational speed. The second case study presented in Chapter 6 pertains

to the design optimization of a spoke-type IPM machine characterized with very high

power density for propulsion application in a formula E racing car. Two objectives of

increasing the power density, and minimizing the aggregate losses over the Le Mans

driving cycle were pursued. The optimization results were used for establishing the

performance trade-offs, and identification of the optimal design rules of high-power-

dense spoke-type IPM machines.

As an additional set of contributions, the last chapter of this dissertation was


240

dedicated to the investigation of three different aspects of the design optimization of

electric machines. First, it was demonstrated that when the minimization of the active

material cost is considered, the optimal design rules of PM machines barely change due

to the variation of the commodity price coefficients. In this case, a cost function based

on the weighted masses of the main active components, and a realistic commodity

price variation by a factor of two was considered. In the continuation of this chapter,

a computationally efficient FE-based method for calculation of strand eddy current

power losses in the stator windings of PM machines, and including this loss component

in the design evaluation process was introduced. This method was implemented on an

example 12-slot 10-pole machine configuration to characterize the strand eddy current

losses under various speeds and loading conditions. The comparison of the results

with a time harmonic transient FE model confirmed the accuracy and computational

efficiency of the developed method. Finally, in this chapter, an analytical method for

calculation of the centrifugal forces on the rotor bridges of V-type IPM machines was

described, and its ability in capturing the influence of the design parameters on such

forces was investigated. Although the presented method cannot take into account the

stress concentrations due to the structural details introduced by some of the design

parameters, such as the PM burial depth or width of the q-axis web, it can provide a

first approximation of the magnitude of stress, which is suitable for implementation

in a large-scale design optimization process.


241

8.2 Recommendation for Future Work

There are numerous research studies in the literature which perform a comparison

between various motor topologies and configurations with an aim to identify the

proper design decision for a given application. It is not uncommon to find such

comparisons being performed between alternative design configurations which are

not necessarily optimized a priori for the specific application, thus leaving much to be

desired for a systematic, objective, and evidence-based comparison. The techniques

and the optimization package developed in this dissertation provide a high-fidelity

simulation platform for realizing such a comparison between different PM motor

configurations for various applications. The studies carried out on PM machines

with 48-slot 8-pole and 12-slot 10-pole configurations in Chapters 3 and 7 can be

mentioned as an example of such analysis.

To bypass the need for a coupled thermal-electromagnetic optimization, and at the

same time meet the thermal performance requirements, the three following measures

were adopted in the developed design optimization package. First, the starting point

of the process was based on an already existing design which meets the thermal

performance requirements. Second, the current density was fixed for all the design

candidates in a given optimization problem with reference to the original design. And

third, the minimization of power losses has been always designated as an objective

of the design optimization. Correspondingly, not only are the designs with power

losses lower than the original design expected to meet the thermal performance

requirements, but the ampere loading of such designs can be further increased, with
242

reference to a post-design optimization thermal performance analysis and as long

as other performance constraints are not violated. This can translate into further

improvement of the optimized designs. Nevertheless, if the computational complexity

of the design optimization is not of concern, the coupled thermal-electromagnetic

approach can be pursued by inspecting the peak and continuous power operation of

the design candidates corresponding to the specifications of the cooling system.

There are other areas of opportunity for improving the developed design

optimization package. To name a few, the implemented cost model can be improved

to also account for the manufacturing cost associated with the power electronics

drives. The loss model can be upgraded to include the additional losses introduced

by the time harmonics of the PWM drives which might be significant during the

extended speed operation. Furthermore, in case the CE-FEA modeling approach

is to be improved, possible methods for taking into account the rotor core and the

magnet losses can be studied and incorporated into the CE-FEA model.
243

BIBLOGRAPHY

[1] M. Lowe, R. Golini, and G. Gereffi, “U.S. Adoption of High-Efficiency Motors


and Drives: Lessons Learned,” Center on Globalization Governance and
Competitiveness, Durham, NC 27705, Tech. Rep., February 2010. [Online].
Available: http://www.cggc.duke.edu

[2] G. A. McCoy and J. G. Douglass, Premium Efficiency Motor Selection and


Application Guide. U.S. Department of Energy’s Office of Energy Efficiency
and Renewable Energy, 2014.

[3] “EV Everywhere Grand Challenge Blueprint,” U.S. Department of Energy,


Washington, DC 20585-0121, Tech. Rep., January 2013. [Online]. Available:
http://www.energy.gov

[4] S. Zhang, J. Xu, J. Junak, D. Fiederling, G. Sawczuk, M. Koch, A. Schalja,


M. Podack, and J. Baumgartner, “Permanent magnet technology for electric
motors in automotive applications,” in Electric Drives Production Conference
(EDPC), 2012 2nd International, 2012, pp. 1–11.

[5] T. Vaimann, A. Kallaste, A. Kilk, and A. Belahcen, “Magnetic properties of


reduced Dy NdFeB permanent magnets and their usage in electrical machines,”
in AFRICON, 2013, 2013, pp. 1–5.

[6] “United State Industrial Electric Motor Systems Market Opportunities


Assessment,” U.S. Department of Energy’s Office of Energy Efficiency and
Renewable Energy, Washington, DC 20585-0121, Tech. Rep., December 2002.
[Online]. Available: http://www.energy.gov

[7] A. EL-Refaie and F. Johnson, “Scalable, Low-Cost, High Performance IPM


Motor for Hybrid Vehicle,” GE Global Research, Niskayuna, NY, Tech. Rep.
DE-FC26-07NT43122, May 2011. [Online]. Available: http://www.energy.gov

[8] W. Larsen, O. Howlett, K. Forsman, and M. Zeller, “Exploring the Customer


Benefits of Permanent Magnet Motors: Test Results and Opportunities for
Next Generation Motor Programs ,” American Council for an Energy-Efficient
Economy, Washington, DC, Tech. Rep., August 2015. [Online]. Available:
http://www.aceee.org

[9] T. M. J.R. Hendershot, Design of brushless permanent-magnet machines.


Motor Design Books LLC; Second Edition, 2010.
244

[10] C. W. Trowbridge and J. K. Sykulski, “Some key developments in


computational electromagnetics and their attribution,” IEEE Transactions on
Magnetics, vol. 42, no. 4, pp. 503–508, 2006.

[11] F. C. Trutt, Analysis of homopolar inductor alternators. University of Delware,


1962.

[12] S. V. Ahamed and E. A. Erdelyi, “Non-Linear Vector Potential Equations


for Highly Saturated Heteropolar Electrical Machines,” IEEE Transactions on
Aerospace, vol. 2, no. 2, pp. 896–903, 1964.

[13] E. A. Erdelyi, S. V. Ahamed, and R. E. Hopkins, “Nonlinear Theory of


Synchronous Machines On-Load,” IEEE Transactions on Power Apparatus and
Systems, vol. PAS-85, no. 7, pp. 792–801, 1966.

[14] S. V. Ahamed and E. A. Erdelyi, “Flux Distribution in DC Machines On-Load


and Overloads,” IEEE Transactions on Power Apparatus and Systems, vol.
PAS-85, no. 9, pp. 960–967, 1966.

[15] E. A. Erdelyi and E. F. Fuchs, “Nonlinear Magnetic Field Analysis of DC


Machines, Part I: Theoretical Fundamentals,” IEEE Transactions on Power
Apparatus and Systems, vol. PAS-89, no. 7, pp. 1546–1554, 1970.

[16] N. A. Demerdash and H. Hamilton, “Effect of Rotor Asymmetry on Field Forms


and Eddy Current Losses in Stator Conductors Due to Radial Flux,” Power
Apparatus and Systems, IEEE Transactions on, vol. PAS-91, no. 5, pp. 1999–
2010, 1972.

[17] N. A. Demerdash, H. B. Hamilton, and G. W. Brown, “Simulation for


Design Purposes of Magnetic Fields in Turbogenerators with Symmetrical and
Asymmetrical Rotors Part I-Model Development and Solution Technique,”
IEEE Transactions on Power Apparatus and Systems, vol. PAS-91, no. 5, pp.
1985–1992, 1972.

[18] N. A. Demerdash and H. B. Hamilton, “Simulation for Design Purposes


of Magnetic Fields in Turbogenerators with Asymmetrical and Symmetrical
Rotors Part II - Model Calibration and Applications,” IEEE Transactions on
Power Apparatus and Systems, vol. PAS-91, no. 5, pp. 1992–1999, 1972.

[19] N. Demerdash and H. Hamilton, “Use of computerized magnetic field solutions


in design optimization of turbogenerators,” IEEE Transactions on Magnetics,
vol. 11, no. 5, pp. 1532–1534, 1975.

[20] N. Demerdash and T. Nehl, “Flexibility and economics of implementation of the


finite element and difference techniques in nonlinear magnetic fields of power
245

devices,” IEEE Transactions on Magnetics, vol. 12, no. 6, pp. 1036–1038, 1976.

[21] ——, “An Evaluation of the Methods of Finite Elements and Finite Differences
in the Solution of Nonlinear Electromagnetic Fields in Electrical Machines,”
IEEE Transactions on Power Apparatus and Systems, vol. PAS-98, no. 1, pp.
74–87, 1979.

[22] P. Silvester and M. V. K. Chari, “Finite Element Solution of Saturable Magnetic


Field Problems,” IEEE Transactions on Power Apparatus and Systems, vol.
PAS-89, no. 7, pp. 1642–1651, 1970.

[23] M. V. K. Chari and P. Silvester, “Finite-Element Analysis of Magnetically


Saturated D-C Machines,” IEEE Transactions on Power Apparatus and
Systems, vol. PAS-90, no. 5, pp. 2362–2372, 1971.

[24] ——, “Analysis of Turboalternator Magnetic Fields by Finite Elements,” IEEE


Transactions on Power Apparatus and Systems, vol. PAS-90, no. 2, pp. 454–464,
1971.

[25] P. Silvester, H. S. Cabayan, and B. T. Browne, “Efficient Techniques for


Finite Element Analysis of Electric Machines,” IEEE Transactions on Power
Apparatus and Systems, vol. PAS-92, no. 4, pp. 1274–1281, 1973.

[26] M. Chari, “Nonlinear finite element solution of electrical machines under no-
load and full-load conditions,” IEEE Transactions on Magnetics, vol. 10, no. 3,
pp. 686–689, 1974.

[27] J. R. Brauer, “Saturated Magnetic Energy Functional for Finite Element


Analysis of Electric Machines,” in IEEE-PES Winter Meeting, 1975.

[28] O. W. Andersen, “Transformer Leakage Flux Program Based on the Finite


Element Method,” IEEE Transactions on Power Apparatus and Systems, vol.
PAS-92, no. 2, pp. 682–689, 1973.

[29] N. Demerdash, T. Nehl, and F. Fouad, “Finite element formulation and


analysis of three dimensional magnetic field problems,” IEEE Transactions on
Magnetics, vol. 16, no. 5, pp. 1092–1094, 1980.

[30] T. Nehl and N. Demerdash, “Application of finite element eddy current analysis
to nondestructive detection of flaws in metallic structures,” IEEE Transactions
on Magnetics, vol. 16, no. 5, pp. 1080–1082, 1980.

[31] N. A. Demerdash and T. W. Nehl, “Dynamic Modeling of Brushless dc Motors


for Aerospace Actuation,” IEEE Transactions on Aerospace and Electronic
Systems, vol. AES-16, no. 6, pp. 811–821, 1980.
246

[32] S. Salon and B. Istfan, “Inverse non-linear finite element problems,” IEEE
Transactions on Magnetics, vol. 22, no. 5, pp. 817–818, 1986.

[33] B. Istfan and S. J. Salon, “Inverse nonlinear finite element problems with local
and global constraints,” IEEE Transactions on Magnetics, vol. 24, no. 6, pp.
2568–2572, 1988.

[34] T. Nakata and N. Takahashi, “Direct finite element analysis of flux and current
distributions under specified conditions,” IEEE Transactions on Magnetics,
vol. 18, no. 2, pp. 325–330, 1982.

[35] ——, “Application of the finite element method to the design of permanent
magnets,” IEEE Transactions on Magnetics, vol. 18, no. 6, pp. 1049–1051,
1982.

[36] ——, “New design method of permanent magnets by using the finite element
method,” IEEE Transactions on Magnetics, vol. 19, no. 6, pp. 2494–2497, 1983.

[37] N. Takahashi, T. Nakata, and N. Uchiyama, “Optimal design method of 3-D


nonlinear magnetic circuit by using magnetization integral equation method,”
IEEE Transactions on Magnetics, vol. 25, no. 5, pp. 4144–4146, 1989.

[38] O. Pironneau, Optimal Shape Design for Elliptic Systems, ser. Springer
series in computational physics. Springer-Verlag, 1984. [Online]. Available:
https://books.google.com/books?id=6JqyAAAAIAAJ

[39] K. Preis and A. Ziegler, “Optimal design of electromagnetic devices with


evolution strategies,” COMPEL, vol. 9, no. A, pp. 119–122, 1990.

[40] S. Subramaniam, S. Kanaganathan, and S. R. H. Hoole, “Two requisite tools


in the optimal design of electromagnetic devices,” IEEE Transactions on
Magnetics, vol. 27, no. 5, pp. 4105–4109, 1991.

[41] M. Guamieri, A. Stella, and F. Trevisan, “A methodological analysis of different


formulations for solving inverse electromagnetic problems,” IEEE Transactions
on Magnetics, vol. 26, no. 2, pp. 622–625, 1990.

[42] A. G. Armstrong, M. Fan, J. Simkin, and C. Trowbridge, “Automated


optimization of magnet design using the boundary integral method,” IEEE
Transactions on Magnetics, vol. 18, no. 2, pp. 620–623, 1982.

[43] S. Gitosusastro, J. L. Coulomb, and J. C. Sabonnadiere, “Performance


derivative calculations and optimization process,” IEEE Transactions on
Magnetics, vol. 25, no. 4, pp. 2834–2839, 1989.
247

[44] C. S. Koh, H. S. Yoon, K. W. Nam, and H. S. Choi, “Magnetic pole shape


optimization of permanent magnet motor for reduction of cogging torque,”
IEEE Transactions on Magnetics, vol. 33, no. 2, pp. 1822–1827, 1997.

[45] I.-H. Park, B.-T. Lee, and S.-Y. Hahn, “Design sensitivity analysis for nonlinear
magnetostatic problems using finite element method,” IEEE Transactions on
Magnetics, vol. 28, no. 2, pp. 1533–1536, 1992.

[46] S. R. H. Hoole, S. Subramaniam, R. Saldanha, J. L. Coulomb, and J. C.


Sabonnadiere, “Inverse problem methodology and finite elements in the
identification of cracks, sources, materials, and their geometry in inaccessible
locations,” IEEE Transactions on Magnetics, vol. 27, no. 3, pp. 3433–3443,
1991.

[47] P. G. Alotto, C. Eranda, B. Brandstatter, G. Furntratt, C. Magele, G. Molinari,


M. Nervi, K. Preis, M. Repetto, and K. R. Richter, “Stochastic algorithms in
electromagnetic optimization,” IEEE Transactions on Magnetics, vol. 34, no. 5,
pp. 3674–3684, 1998.

[48] R. R. Saldanha, J. L. Coulomb, A. Foggia, and J. C. Sabonnadiere, “A dual


method for constrained optimization design in magnetostatic problems,” IEEE
Transactions on Magnetics, vol. 27, no. 5, pp. 4136–4141, 1991.

[49] R. R. Saldanha, J. L. Coulomb, and J. C. Sabonnadiere, “An ellipsoid algorithm


for the optimum design of magnetostatic problems,” IEEE Transactions on
Magnetics, vol. 28, no. 2, pp. 1573–1576, 1992.

[50] K. Weeber and S. R. H. Hoole, “Structural design optimization as a technology


source for developments in the electromagnetics domain,” IEEE Transactions
on Magnetics, vol. 29, no. 2, pp. 1807–1811, 1993.

[51] A. Gottvald, “Comparative analysis of optimization methods for magnetostat-


ics,” IEEE Transactions on Magnetics, vol. 24, no. 1, pp. 411–414, 1988.

[52] ——, “Optimal magnet design for NMR,” IEEE Transactions on Magnetics,
vol. 26, no. 2, pp. 399–402, 1990.

[53] K. Preis, O. Biro, M. Friedrich, A. Gottvald, and C. Magele, “Comparison of


different optimization strategies in the design of electromagnetic devices,” IEEE
Transactions on Magnetics, vol. 27, no. 5, pp. 4154–4157, 1991.

[54] A. Gottvald, K. Preis, C. Magele, O. Biro, and A. Savini, “Global


optimization methods for computational electromagnetics,” IEEE Transactions
on Magnetics, vol. 28, no. 2, pp. 1537–1540, 1992.
248

[55] C. A. Magele, K. Preis, W. Renhart, R. Dyczij-Edlinger, and K. R. Richter,


“Higher order evolution strategies for the global optimization of electromagnetic
devices,” IEEE Transactions on Magnetics, vol. 29, no. 2, pp. 1775–1778, 1993.

[56] G. Drago, A. Manella, M. Nervi, M. Repetto, and G. Secondo, “A combined


strategy for optimization in nonlinear magnetic problems using simulated
annealing and search techniques,” IEEE Transactions on Magnetics, vol. 28,
no. 2, pp. 1541–1544, 1992.

[57] J. Simkin and C. W. Trowbridge, “Optimizing electromagnetic devices


combining direct search methods with simulated annealing,” IEEE Transactions
on Magnetics, vol. 28, no. 2, pp. 1545–1548, 1992.

[58] S. Russenschuck, “Mathematical optimization techniques for the design of


permanent magnet synchronous machines based on numerical field calculation,”
IEEE Transactions on Magnetics, vol. 26, no. 2, pp. 638–641, 1990.

[59] S. Ratnajeevan, H. Hoole, K. Weeber, and S. Subramaniam, “Fictitious minima


of object functions, finite element meshes, and edge elements in electromagnetic
device synthesis,” IEEE Transactions on Magnetics, vol. 27, no. 6, pp. 5214–
5216, 1991.

[60] K. Weeber and S. R. H. Hoole, “A structural mapping technique for geometric


parametriiation in the synthesis of magnetic devices,” Int. J. Num. Meth. Eng.,
vol. 33, pp. 2145–2179, 1992.

[61] ——, “Geometric parametrization and constrained optimization techniques


in the design of salient pole synchronous machines,” IEEE Transactions on
Magnetics, vol. 28, no. 4, pp. 1948–1960, 1992.

[62] K. Preis, C. Magele, and O. Biro, “FEM and evolution strategies in the optimal
design of electromagnetic devices,” IEEE Transactions on Magnetics, vol. 26,
no. 5, pp. 2181–2183, 1990.

[63] M. Schafer-Jotter and W. Muller, “Optimization of electrotechnical devices


using a numerical laboratory,” IEEE Transactions on Magnetics, vol. 26, no. 2,
pp. 815–818, 1990.

[64] J. Simkin and C. W. Trowbridge, “Optimization problems in electromagnetics,”


IEEE Transactions on Magnetics, vol. 27, no. 5, pp. 4016–4019, 1991.

[65] G. F. Uler, O. A. Mohammed, and C.-S. Koh, “Utilizing genetic algorithms


for the optimal design of electromagnetic devices,” IEEE Transactions on
Magnetics, vol. 30, no. 6, pp. 4296–4298, 1994.
249

[66] O. A. Mohammed, “Practical issues in the application of genetic algorithms


to optimal design problems in electromagnetics,” in Southeastcon ’96. Bringing
Together Education, Science and Technology., Proceedings of the IEEE, 1996,
pp. 634–640.

[67] J.-S. Chun, H.-K. Jung, and J.-S. Yoon, “Shape optimization of closed slot
type permanent magnet motors for cogging torque reduction using evolution
strategy,” IEEE Transactions on Magnetics, vol. 33, no. 2, pp. 1912–1915,
1997.

[68] T. K. Chung, S. K. Kim, and S.-Y. Hahn, “Optimal pole shape design for the
reduction of cogging torque of brushless DC motor using evolution strategy,”
IEEE Transactions on Magnetics, vol. 33, no. 2, pp. 1908–1911, 1997.

[69] N. Bianchi and S. Bolognani, “Design optimisation of electric motors by genetic


algorithms,” IEE Proceedings - Electric Power Applications, vol. 145, no. 5, pp.
475–483, 1998.

[70] N. Bianchi and A. Canova, “FEM analysis and optimisation design of an


IPM synchronous motor,” in Power Electronics, Machines and Drives, 2002.
International Conference on (Conf. Publ. No. 487), 2002, pp. 49–54.

[71] T. Ohnishi and N. Takahashi, “Optimal design of efficient IPM motor using
finite element method,” IEEE Transactions on Magnetics, vol. 36, no. 5, pp.
3537–3539, 2000.

[72] F. A. Fouad, T. W. Nehl, and N. A. Demerdash, “Magnetic Field Modeling of


Permanent Magnet Type Electronically Operated Synchronous Machines Using
Finite Elements,” IEEE Transactions on Power Apparatus and Systems, vol.
PAS-100, no. 9, pp. 4125–4135, 1981.

[73] B. Cassimere and S. Sudhoff, “Population-Based Design of Surface-


Mounted Permanent-Magnet Synchronous Machines,” Energy Conversion,
IEEE Transactions on, vol. 24, no. 2, pp. 338–346, June 2009.

[74] E. C. Lovelace, T. M. Jahns, and J. H. Lang, “Impact of saturation and


inverter cost on interior PM synchronous machine drive optimization,” IEEE
Transactions on Industry Applications, vol. 36, no. 3, pp. 723–729, 2000.

[75] A. A. Arkadan and Y. Chen, “Artificial neural network for the inverse
electromagnetic problem of system identification,” in Southeastcon ’94. Creative
Technology Transfer - A Global Affair., Proceedings of the 1994 IEEE, 1994, pp.
162–164.

[76] G. Y. Sizov, D. M. Ionel, and N. A. O. Demerdash, “A review of efficient FE


250

modeling techniques with applications to PM AC machines,” in 2011 IEEE


Power and Energy Society General Meeting, 2011, pp. 1–6.

[77] Y. Duan and D. M. Ionel, “A Review of Recent Developments in Electrical


Machine Design Optimization Methods With a Permanent-Magnet Synchronous
Motor Benchmark Study,” IEEE Transactions on Industry Applications, vol. 49,
no. 3, pp. 1268–1275, 2013.

[78] G. Bramerdorfer, A. Zavoianu, S. Silber, E. Lughofer, and W. Amrhein, “Speed


Improvements for the Optimization of Electrical Machines - a Survey,” in
Electric Machines Drives Conference (IEMDC), 2015 IEEE International, May
2015.

[79] F. Poltschak and W. Amrhein, “A dynamic nonlinear model for permanent


magnet synchronous machines,” in 2008 IEEE International Symposium on
Industrial Electronics, 2008, pp. 724–729.

[80] D. M. Ionel and M. Popescu, “Finite-Element Surrogate Model for


Electric Machines with Revolving Field-Application to IPM Motors,” IEEE
Transactions on Industry Applications, vol. 46, no. 6, pp. 2424 –2433, Nov.-
Dec. 2010.

[81] D. M. Ionel and M. M. Popescu, “Ultrafast Finite-Element Analysis of Brushless


PM Machines Based on SpaceTime Transformations,” IEEE Transactions on
Industry Applications, vol. 47, no. 2, pp. 744 –753, March-April 2011.

[82] G. Y. Sizov, D. M. Ionel, and N. A. O. Demerdash, “Modeling and Parametric


Design of Permanent-Magnet AC Machines Using Computationally Efficient
Finite-Element Analysis,” IEEE Transactions on Industrial Electronics, vol. 59,
no. 6, pp. 2403–2413, 2012.

[83] W. Jiang, T. M. Jahns, T. A. Lipo, W. Taylor, and Y. Suzuki, “Machine design


optimization based on finite element analysis in a high-throughput computing
environment,” in 2012 IEEE Energy Conversion Congress and Exposition
(ECCE), 2012, pp. 869–876.

[84] G. Sizov, P. Zhang, D. Ionel, N. Demerdash, and M. Rosu, “Automated Multi-


Objective Design Optimization of PM AC Machines Using Computationally
Efficient- FEA and Differential Evolution,” IEEE Transactions on Industry
Applications, vol. 49, no. 5, pp. 2086–2096, 2013.

[85] P. Zhang, G. Sizov, M. Li, D. Ionel, N. Demerdash, S. Stretz, and


A. Yeadon, “Multi-Objective Tradeoffs in the Design Optimization of a
Brushless Permanent-Magnet Machine With Fractional-Slot Concentrated
Windings,” Industry Applications, IEEE Transactions on, vol. 50, no. 5, pp.
251

3285–3294, 2014.

[86] I. P. Brown, M. W. Critchley, J. Yin, S. B. Memory, G. Y. Sizov, S. W. Elbel,


C. D. Bowers, M. Petersen, and P. S. Hrnjak, “Design and Evaluation of Interior
Permanent-Magnet Compressor Motors for Commercial Transcritical CO2 (R-
744) Heat Pump Water Heaters,” IEEE Transactions on Industry Applications,
vol. 51, no. 1, pp. 576–586, 2015.

[87] Y. Duan and D. Ionel, “Nonlinear Scaling Rules for Brushless PM Synchronous
Machines Based on Optimal Design Studies for a Wide Range of Power
Ratings,” Industry Applications, IEEE Transactions on, vol. 50, no. 2, pp.
1044–1052, March 2014.

[88] P. Zhang, G. Sizov, D. Ionel, and N. Demerdash, “Establishing the Relative


Merits of Interior and Spoke-Type Permanent-Magnet Machines With Ferrite
or NdFeB Through Systematic Design Optimization,” Industry Applications,
IEEE Transactions on, vol. 51, no. 4, pp. 2940–2948, 2015.

[89] Y. Wang, D. Ionel, and D. Staton, “Ultrafast Steady-state Multi-physics Model


for PM and Synchronous Reluctance Machines,” Industry Applications, IEEE
Transactions on, vol. PP, no. 99, pp. 1–1, 2015.

[90] Y. Wang, D. M. Ionel, M. Jiang, and S. J. Stretz, “Establishing the Relative


Merits of Synchronous Reluctance and PM Assisted Technology Through
Systematic Design Optimization,” IEEE Transactions on Industry Applications,
vol. PP, no. 99, pp. 1–1, 2016.

[91] S. E. Sibande, M. J. Kamper, R. Wang, and E. T. Rakgati, “Optimal design of a


PM-assisted rotor of a 110 kW reluctance synchronous machine,” in AFRICON,
2004. 7th AFRICON Conference in Africa, vol. 2, 2004, pp. 793–797 Vol.2.

[92] D. Zarko, D. Ban, and T. A. Lipo, “Design optimization of interior permanent


magnet (IPM) motors with maximized torque output in the entire speed range,”
in Power Electronics and Applications, 2005 European Conference on, 2005, pp.
10 pp.–P.10.

[93] R. Schiferl and T. Lipo, “Power capability of salient pole permanent magnet
synchronous motors in variable speed drive applications,” Industry Applications,
IEEE Transactions on, vol. 26, no. 1, pp. 115–123, Jan 1990.

[94] W. Ouyang, D. Zarko, and T. Lipo, “Permanent Magnet Machine Design


Practice and Optimization,” in Industry Applications Conference, 2006. 41st
IAS Annual Meeting. Conference Record of the 2006 IEEE, vol. 4, 2006, pp.
1905–1911.
252

[95] G. Pellegrino and F. Cupertino, “IPM motor rotor design by means of FEA-
based multi-objective optimization,” in Industrial Electronics (ISIE), 2010
IEEE International Symposium on, 2010, pp. 1340–1346.

[96] P. Zhang, D. Ionel, and N. Demerdash, “Saliency ratio and power factor of
IPM motors optimally designed for high efficiency and low cost objectives,” in
Energy Conversion Congress and Exposition (ECCE), 2014 IEEE, Sept 2014,
pp. 3541–3547.

[97] S. Morimoto, S. Ooi, Y. Inoue, and M. Sanada, “Experimental Evaluation


of a Rare-Earth-Free PMASynRM With Ferrite Magnets for Automotive
Applications,” Industrial Electronics, IEEE Transactions on, vol. 61, no. 10,
pp. 5749–5756, 2014.

[98] A. Fatemi, N. Demerdash, and D. Ionel, “Design optimization of IPM


machines for efficient operation in extended speed range,” in Transportation
Electrification Conference and Expo (ITEC), 2015 IEEE, 2015, pp. 1–8.

[99] F. Parasiliti, M. Villani, S. Lucidi, and F. Rinaldi, “Finite-Element-Based


Multiobjective Design Optimization Procedure of Interior Permanent Magnet
Synchronous Motors for Wide Constant-Power Region Operation,” IEEE
Transactions on Industrial Electronics, vol. 59, no. 6, pp. 2503 –2514, June
2012.

[100] J. hee Lee and B.-I. Kwon, “Optimal Rotor Shape Design of a Concentrated
Flux IPM-Type Motor for Improving Efficiency and Operation Range,”
Magnetics, IEEE Transactions on, vol. 49, no. 5, pp. 2205–2208, 2013.

[101] K. Yamazaki, M. Kumagai, T. Ikemi, and S. Ohki, “A Novel Rotor Design of


Interior Permanent-Magnet Synchronous Motors to Cope with Both Maximum
Torque and Iron-Loss Reduction,” Industry Applications, IEEE Transactions
on, vol. 49, no. 6, pp. 2478–2486, 2013.

[102] K. Yamazaki and H. Ishigami, “Rotor-Shape Optimization of Interior-


Permanent-Magnet Motors to Reduce Harmonic Iron Losses,” IEEE Trans-
actions on Industrial Electronics, vol. 57, no. 1, pp. 61–69, 2010.

[103] F. Cupertino, G. Pellegrino, E. Armando, and C. Gerada, “A SyR and IPM


machine design methodology assisted by optimization algorithms,” in Energy
Conversion Congress and Exposition (ECCE), 2012 IEEE, 2012, pp. 3686–3691.

[104] M. Barcaro, N. Bianchi, and F. Magnussen, “Permanent-Magnet Optimization


in Permanent-Magnet-Assisted Synchronous Reluctance Motor for a Wide
Constant-Power Speed Range,” Industrial Electronics, IEEE Transactions on,
vol. 59, no. 6, pp. 2495–2502, 2012.
253

[105] P. Lazari, J. Wang, and L. Chen, “A Computationally Efficient Design


Technique for Electric-Vehicle Traction Machines,” Industry Applications, IEEE
Transactions on, vol. 50, no. 5, pp. 3203–3213, Sept 2014.

[106] J. Wang, X. Yuan, and K. Atallah, “Design Optimization of a Surface-Mounted


Permanent-Magnet Motor With Concentrated Windings for Electric Vehicle
Applications,” Vehicular Technology, IEEE Transactions on, vol. 62, no. 3, pp.
1053–1064, March 2013.

[107] E. Carraro, M. Morandin, and N. Bianchi, “Traction PMASR Motor


Optimization According to a Given Driving Cycle,” IEEE Transactions on
Industry Applications, vol. 52, no. 1, pp. 209–216, 2016.

[108] A. Fatemi, D. M. Ionel, N. A. O. Demerdash, and T. W. Nehl, “Optimal Design


of IPM Motors with Different Cooling Systems and Winding Configurations,”
IEEE Transactions on Industry Applications, vol. PP, no. 99, pp. 1–1, 2016.

[109] A. Fatemi, N. Demerdash, T. Nehl, and D. Ionel, “Large-scale Design


Optimization of PM Machines Over a Target Operating Cycle,” IEEE
Transactions on Industry Applications, vol. PP, no. 99, pp. 1–1, 2016.

[110] A. Fatemi, D. M. Ionel, N. A. O. Demerdash, and T. W. Nehl, “Fast


Multi-Objective CMODE-Type Optimization of PM Machines Using Multicore
Desktop Computers,” IEEE Transactions on Industry Applications, vol. PP,
no. 99, pp. 1–1, 2016.

[111] A. Fatemi, D. M. Ionel, M. Popescu, and N. A. O. Demerdash, “Design


optimization of spoke-type pm motors for formula e racing cars,” in 2016 IEEE
Energy Conversion Congress and Exposition (ECCE), 2016, p. in press.

[112] A. Fatemi, D. M. Ionel, N. A. O. Demerdash, S. Stretz, and T. M. Jahns,


“RSM-DE-ANN Method for Sensitivity Analysis of Active Material Cost in PM
Motors,” in 2016 IEEE Energy Conversion Congress and Exposition (ECCE),
2016, p. in press.

[113] A. Fatemi, D. M. Ionel, N. A. O. Demerdash, D. Staton, R. Wrobel, and


Y. C. Chong, “Computationally Efficient Method for Calculation of Strand
Eddy Current Losses in Stator Windings of Electric Machines,” in 2016 IEEE
Energy Conversion Congress and Exposition (ECCE), 2016, pp. 4383–4390.

[114] A. Fatemi, N. A. O. Demerdash, D. M. Ionel, and T. W. Nehl, “Large-scale


electromagnetic design optimization of PM machines over a target operating
cycle,” in 2015 IEEE Energy Conversion Congress and Exposition (ECCE),
2015, pp. 4383–4390.
254

[115] A. Fatemi, D. M. Ionel, N. A. O. Demerdash, and T. W. Nehl, “Fast


multi-objective CMODE-type optimization of electric machines for multicore
desktop computers,” in 2015 IEEE Energy Conversion Congress and Exposition
(ECCE), 2015, pp. 5593–5600.

[116] A. Fatemi, D. Ionel, and N. Demerdash, “Identification of design rules for


interior PM motors with different cooling systems,” in Electric Machines Drives
Conference (IEMDC), 2015 IEEE International, May 2015.

[117] G. Y. Sizov, Design synthesis and optimization of permanent magnet


synchronous machines based on computationally-efficient finite element
analysis. Marquette University, 2013.

[118] P. Zhang, A novel design optimization of a fault-tolerant ac permanent magnet


machine-drive system. Marquette University, 2013.

[119] G. Y. Sizov, D. M. Ionel, and N. A. O. Demerdash, “Modeling and Parametric


Design of Permanent-Magnet AC Machines Using Computationally Efficient-
Finite Element Analysis,” IEEE Transactions on Industrial Electronics, vol. 59,
no. 6, pp. 2403 –2413, June 2012.

[120] K. V. Price, R. M. Storn, and J. A. Lampinen, Differential Evolution-A Practical


Approach to Global Optimization. Springer-Verlag Berlin Heidelberg, 2005.

[121] J. Coulomb and G. Meunier, “Finite element implementation of virtual


work principle for magnetic or electric force and torque computation,” IEEE
Transactions on Magnetics, vol. 20, no. 5, pp. 1894–1896, 1984.

[122] D. Ionel, M. Popescu, M. McGilp, T. Miller, S. Dellinger, and R. Heideman,


“Computation of Core Losses in Electrical Machines Using Improved Models for
Laminated Steel,” IEEE Transactions on Industry Applications, vol. 43, no. 6,
pp. 1554 –1564, nov.-dec. 2007.

[123] A. Boglietti, A. Cavagnino, D. Ionel, M. Popescu, D. Staton, and S. Vaschetto,


“A General Model to Predict the Iron Losses in PWM Inverter-Fed Induction
Motors,” IEEE Transactions on Industry Applications, vol. 46, no. 5, pp. 1882–
1890, 2010.

[124] D. Ionel, M. Popescu, C. Cossar, M. McGilp, A. Boglietti, and A. Cavagnino,


“A General Model of the Laminated Steel Losses in Electric Motors with PWM
Voltage Supply,” in IEEE Industry Applications Society Annual Meeting (IAS),
2008, pp. 1–7.

[125] M. Popescu, D. Ionel, A. Boglietti, A. Cavagnino, C. Cossar, and M. McGilp,


“A General Model for Estimating the Laminated Steel Losses Under PWM
255

Voltage Supply,” IEEE Transactions on Industry Applications, vol. 46, no. 4,


pp. 1389–1396, 2010.

[126] A. EL-Refaie, “Fractional-Slot Concentrated-Windings Synchronous Permanent


Magnet Machines: Opportunities and Challenges,” IEEE Transactions on
Industrial Electronics, vol. 57, no. 1, pp. 107–121, 2010.

[127] K. Zielinski, P. Weitkemper, R. Laur, and R. L. K.-D. Kammeyer, K. Zielinski,


“Examination of Stopping Criteria for Differential Evolution based on a Power
Allocation Problem,” in 10th International Conference on Optimization of
Electrical and Electronic Equipment, Brasov, Romania, May 2006.

[128] K. Zielinski and R. Laur, “Stopping Criteria for a Constrained Single-Objective


Particle Swarm Optimization Algorithm,” Informatica (03505596), vol. 31, pp.
51–59, March 2007.

[129] N. Bianchi and S. Bolognani, “Brushless DC motor design: an optimisation


procedure based on genetic algorithms,” in Electrical Machines and Drives,
1997 Eighth International Conference on (Conf. Publ. No. 444), Sep 1997, pp.
16–20.

[130] M. Olszewski, “Evaluation of 2004 Toyota Prius Hybrid Electric Drive System,”
in Oak Ridge National Laboratory, U.S. Department of Energy, Sept 2004.

[131] ——, “Evaluation of the 2007 Toyota Camry Hybrid Synergy Drive System,”
in Oak Ridge National Laboratory, U.S. Department of Energy, Apr 2008.

[132] D. Dorrell, M. Hsieh, M. Popescu, L. Evans, D. Staton, and V. Grout, “A


Review of the Design Issues and Techniques for Radial-Flux Brushless Surface
and Internal Rare-Earth Permanent-Magnet Motors,” Industrial Electronics,
IEEE Transactions on, vol. 58, no. 9, pp. 3741–3757, 2011.

[133] E. Armando, P. Guglielmi, G. Pellegrino, M. Pastorelli, and A. Vagati,


“Accurate Modeling and Performance Analysis of IPM-PMASR Motors,”
Industry Applications, IEEE Transactions on, vol. 45, no. 1, pp. 123–130, 2009.

[134] W. Jiang and T. Jahns, “Coupled electromagnetic/thermal machine design


optimization based on finite element analysis with application of artificial neural
network,” in Energy Conversion Congress and Exposition (ECCE), 2014 IEEE,
Sept 2014, pp. 5160–5167.

[135] S. Semidey, Y. Duan, J. Mayor, R. Harley, and T. Habetler, “Optimal


Electromagnetic-Thermo-Mechanical Integrated Design Candidate Search and
Selection for Surface-Mount Permanent-Magnet Machines Considering Load
256

Profiles,” Industry Applications, IEEE Transactions on, vol. 47, no. 6, pp. 2460–
2468, 2011.

[136] J. Wang and D. Howe, “Design optimization of radially magnetized, iron-cored,


tubular permanent-magnet machines and drive systems,” Magnetics, IEEE
Transactions on, vol. 40, no. 5, pp. 3262–3277, 2004.

[137] G. Sizov, D. Ionel, and N. Demerdash, “Multi-Objective Optimization of


PM AC Machines Using Computationally Efficient-FEA and Differential
Evolution,” in IEEE International Electric Machines Drives Conference 2011
(IEMDC ’11), may 2011, pp. 1528 –1533.

[138] G. Lei, C. Liu, J. Zhu, and Y. Guo, “Techniques for Multilevel Design
Optimization of Permanent Magnet Motors,” Energy Conversion, IEEE
Transactions on, vol. PP, no. 99, pp. 1–11, 2015.

[139] C. Xia, L. Guo, Z. Zhang, T. Shi, and H. Wang, “Optimal Designing of


Permanent Magnet Cavity to Reduce Iron Loss of Interior Permanent Magnet
Machine,” Magnetics, IEEE Transactions on, vol. 51, no. 12, pp. 1–9, 2015.

[140] F. Parasiliti, M. Villani, S. Lucidi, and F. Rinaldi, “A new optimization


approach for the design of IPM synchronous motor with wide constant-power
region,” in Electrical Machines (ICEM), 2010 XIX International Conference
on, 2010, pp. 1–7.

[141] F. Dubas, C. Espanet, and A. Miraoui, “Design of a high-speed permanent


magnet motor for the drive of a fuel cell air-compressor,” in Vehicle Power and
Propulsion, 2005 IEEE Conference, 2005, pp. 8 pp.–.

[142] M. Popescu, I. Foley, D. A. Staton, and J. E. Goss, “Multi-physics analysis


of a high torque density motor for electric racing cars,” in Energy Conversion
Congress and Exposition, 2015. ECCE 2015. IEEE, Sept 2015.

[143] A. Boglietti, A. Cavagnino, D. Staton, M. Shanel, M. Mueller, and C. Mejuto,


“Evolution and Modern Approaches for Thermal Analysis of Electrical
Machines,” Industrial Electronics, IEEE Transactions on, vol. 56, no. 3, pp.
871–882, March 2009.

[144] Z. Azar, Z. Zhu, and G. Ombach, “Influence of Electric Loading and Magnetic
Saturation on Cogging Torque, Back-EMF and Torque Ripple of PM Machines,”
Magnetics, IEEE Transactions on, vol. 48, no. 10, pp. 2650–2658, 2012.

[145] J. Cros and P. Viarouge, “Synthesis of high performance PM motors with


concentrated windings,” Energy Conversion, IEEE Transactions on, vol. 17,
no. 2, pp. 248–253, 2002.
257

[146] 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, vol. 3, 2004, pp. 1696–1702 vol.3.

[147] K. Hirota, H. Nakamura, T. Minowa, and M. Honshima, “Coercivity


Enhancement by the Grain Boundary Diffusion Process to Nd amp;8211;Fe
amp;8211;B Sintered Magnets,” Magnetics, IEEE Transactions on, vol. 42,
no. 10, pp. 2909–2911, 2006.

[148] S. Jurkovic, K. Rahman, N. Patel, and P. Savagian, “Next Generation Voltec


Electric Machines; Design and Optimization for Performance and Rare-Earth
Mitigation,” SAE Intnational Journal of Alternative Powertrains, vol. 4, no. 2,
pp. 336–342, Apr 2015.

[149] G. Pellegrino and F. Cupertino, “FEA-based Multi-Objective Optimization


of IPM Motor Design Including Rotor Losses,” in IEEE Energy Conversion
Congress and Exposition (ECCE), Sept. 2010, pp. 3659–3666.

[150] R. H. Myers and D. C. Montgomery, Response Surface Methodology: Process


and Product Optimization Using Designed Experiments . John Wiley & Sons,
2002.

[151] Y. Wang and Z. Cai, “Combining Multiobjective Optimization With Differential


Evolution to Solve Constrained Optimization Problems,” Evolutionary
Computation, IEEE Transactions on, vol. 16, no. 1, pp. 117–134, Feb 2012.

[152] R. Ramarathnam, B. Desai, and V. Rao, “A Comparative Study of


Minimization Techniques for Optimization of Induction Motor Design,” Power
Apparatus and Systems, IEEE Transactions on, vol. PAS-92, no. 5, pp. 1448–
1454, Sept 1973.

[153] D.-J. Sim, D.-H. Cho, J.-S. Chun, H.-K. Jung, and T.-K. Chung, “Efficiency
optimization of interior permanent magnet synchronous motor using genetic
algorithms,” Magnetics, IEEE Transactions on, vol. 33, no. 2, pp. 1880–1883,
Mar 1997.

[154] R. O. Kuehl, Design of Experiments: Statistical Principle of Research Design


and Anlysis, Second edition. Duxbury/Thomson Learning, 2000.

[155] D. Zarko and S. Stipetic, “Criteria for optimal design of interior permanent
magnet motor series,” in Electrical Machines (ICEM), 2012 XXth International
Conference on, Sept 2012, pp. 1242–1249.

[156] D. Zarko, D. Ban, and D. Goricki, “Improvement of a Servo Motor Design


258

Including Optimization and Cost Analysis,” in 12th International Power


Electronics and Motion Control Conference (EPE-PEMC), 2006, pp. 302–307.

[157] E. Mezura Montes and C. A. Coello Coello, “A simple multimembered


evolution strategy to solve constrained optimization problems,” Evolutionary
Computation, IEEE Transactions on, vol. 9, no. 1, pp. 1–17, Feb 2005.

[158] K. Weeber and S. Hoole, “Geometric parametrization and constrained


optimization techniques in the design of salient pole synchronous machines,”
Magnetics, IEEE Transactions on, vol. 28, no. 4, pp. 1948–1960, Jul 1992.

[159] E. Lovelace, T. Jahns, T. Keim, and J. H. Lang, “Mechanical design con-


siderations for conventionally laminated, high-speed, interior PM synchronous
machine rotors,” Industry Applications, IEEE Transactions on, vol. 40, no. 3,
pp. 806–812, May 2004.

[160] E. Zitzler and L. Thiele, “Multiobjective evolutionary algorithms: a


comparative case study and the strength Pareto approach,” Evolutionary
Computation, IEEE Transactions on, vol. 3, no. 4, pp. 257–271, 1999.

[161] C. M. Fonseca, P. J. Fleming, E. Zitzler, L. Thiele, and K. Deb, “The measure of


Pareto optima: Applications in multiobjective metaheuristics,” in Evolutionary
Multi-Criterion Optimization: Second International Conference, EMO 2003,
2003, pp. 519–533.

[162] P. Zhang, G. Sizov, J. He, D. Ionel, and N. Demerdash, “Calculation


of Magnet Losses in Concentrated-Winding Permanent-Magnet Synchronous
Machines Using a Computationally Efficient Finite-Element Method,” Industry
Applications, IEEE Transactions on, vol. 49, no. 6, pp. 2524–2532, 2013.

[163] J. Goss, P. Mellor, R. Wrobel, D. Staton, and M. Popescu, “The design of


AC permanent magnet motors for electric vehicles: A computationally efficient
model of the operational envelope,” in Power Electronics, Machines and Drives
(PEMD 2012), 6th IET International Conference on, 2012, pp. 1–6.

[164] J. Goss, R. Wrobel, P. Mellor, and D. Staton, “The design of AC permanent


magnet motors for electric vehicles: A design methodology,” in Electric
Machines Drives Conference (IEMDC), 2013 IEEE International, 2013, pp.
871–878.

[165] T. Jahns, “Flux-Weakening Regime Operation of an Interior Permanent-Magnet


Synchronous Motor Drive,” Industry Applications, IEEE Transactions on, vol.
IA-23, no. 4, pp. 681–689, July 1987.

[166] E. H. M and J. C. Balda, “Permanent magnet synchronous motor drive for HEV
259

propulsion: optimum speed ratio and parameter determination,” in Vehicular


Technology Conference, 2002. Proceedings. VTC 2002-Fall. 2002 IEEE 56th,
vol. 3, 2002, pp. 1500–1504 vol.3.

[167] Z. Rahman, K. Butler, and M. Ehsani, “Effect of Extended-speed, Constant-


power Operation of Electric Drives on the Design and Performance of EV
Propulsion System,” in SAE Future Car Congress, Paper No. 2001-01-0699,
2000.

[168] S. Z. Selim and M. A. Ismail, “K-Means-Type Algorithms: A Generalized


Convergence Theorem and Characterization of Local Optimality,” Pattern
Analysis and Machine Intelligence, IEEE Transactions on, vol. PAMI-6, no. 1,
pp. 81–87, Jan 1984.

[169] B. Stumberger, G. Stumberger, D. Dolinar, A. Hamler, and M. Trlep,


“Evaluation of saturation and cross-magnetization effects in interior permanent-
magnet synchronous motor,” Industry Applications, IEEE Transactions on,
vol. 39, no. 5, pp. 1264–1271, 2003.

[170] D. Ionel, J. Eastham, T. Miller, and E. Demeter, “Design considerations


for permanent magnet synchronous motors for flux weakening applications,”
Electric Power Applications, IEE Proceedings -, vol. 145, no. 5, pp. 435–440,
Sep 1998.

[171] A. Adnanes, “Torque analysis of permanent magnet synchronous motors,”


in Power Electronics Specialists Conference, 1991. PESC ’91 Record., 22nd
Annual IEEE, Jun 1991, pp. 695–701.

[172] 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, no. 5,
pp. 866–871, Sep 1990.

[173] V. Zivotic-Kukolj, W. Soong, and N. Ertugrul, “Iron Loss Reduction


in an Interior PM Automotive Alternator,” Industry Applications, IEEE
Transactions on, vol. 42, no. 6, pp. 1478–1486, 2006.

[174] B. Chalmers, L. Musaba, and D. Gosden, “Variable-frequency synchronous


motor drives for electric vehicles,” Industry Applications, IEEE Transactions
on, vol. 32, no. 4, pp. 896–903, Jul 1996.

[175] J. Miller, “Oak Ridge National Laboratory Annual Progress Report for the
Power Electronics and Electric Motors Program,” in Oak Ridge National
Laboratory, U.S. Department of Energy, Nov 2013.
260

[176] A. Borisavljevic, H. Polinder, and J. A. Ferreira, “On the Speed Limits of


Permanent-Magnet Machines,” IEEE Transactions on Industrial Electronics,
vol. 57, no. 1, pp. 220–227, 2010.

[177] G.-H. Kang, J. Hur, H.-G. Sung, and J.-P. Hong, “Optimal design of spoke type
BLDC motor considering irreversible demagnetization of permanent magnet,”
in Electrical Machines and Systems, 2003. ICEMS 2003. Sixth International
Conference on, vol. 1, 2003, pp. 234–237 vol.1.

[178] B. kuk Lee, G.-H. Kang, J. Hur, and D.-W. You, “Design of spoke type
BLDC motors with high power density for traction applications,” in Industry
Applications Conference, 2004. 39th IAS Annual Meeting. Conference Record
of the 2004 IEEE, vol. 2, 2004, pp. 1068–1074 vol.2.

[179] E. Carraro, N. Bianchi, S. Zhang, and M. Koch, “Permanent magnet volume


minimization of spoke type fractional slot synchronous motors,” in Energy
Conversion Congress and Exposition (ECCE), 2014 IEEE, 2014, pp. 4180–4187.

[180] H.-W. Kim, K.-T. Kim, Y.-S. Jo, and J. Hur, “Optimization Methods of Torque
Density for Developing the Neodymium Free SPOKE-Type BLDC Motor,”
Magnetics, IEEE Transactions on, vol. 49, no. 5, pp. 2173–2176, 2013.

[181] D. Ionel, D. Jackson, G. Starr, and A. Turner, “Permanent magnet brushless


motors for industrial variable speed drives,” in Power Electronics, Machines
and Drives, 2002. International Conference on (Conf. Publ. No. 487), 2002,
pp. 650–654.

[182] “Critical Materials Strategy,” U.S. Department of Energy’s Office of Energy


Efficiency and Renewable Energy, Washington, DC 20585-0121, Tech. Rep.,
December 2011. [Online]. Available: http://www.energy.gov

[183] M. Barcaro and N. Bianchi, “Interior PM Machines Using Ferrite to Replace


Rare-Earth Surface PM Machines,” Industry Applications, IEEE Transactions
on, vol. 50, no. 2, pp. 979–985, 2014.

[184] I. Petrov and J. Pyrhonen, “Performance of Low-Cost Permanent Magnet


Material in PM Synchronous Machines,” Industrial Electronics, IEEE
Transactions on, vol. 60, no. 6, pp. 2131–2138, 2013.

[185] M. Barcaro, N. Bianchi, and F. Magnussen, “Permanent-Magnet Optimization


in Permanent-Magnet-Assisted Synchronous Reluctance Motor for a Wide
Constant-Power Speed Range,” IEEE Transaction on Industrial Electronics,
vol. 59, no. 6, pp. 2495–2502, June 2012.

[186] E. C. Lovelace, T. M. Jahns, J. L. Kirtley, and J. H. Lang, “An interior PM


261

starter/alternator for automotive applications,” in Proc. ICEM, vol. 3, 1998,


pp. 1802–1808.

[187] G. Lei, C. Liu, Y. Guo, and J. Zhu, “Robust Multidisciplinary Design


Optimization of PM Machines With Soft Magnetic Composite Cores for Batch
Production,” IEEE Transactions on Magnetics, vol. 52, no. 3, pp. 1–4, 2016.

[188] M. Hagan, H. Demuth, M. Beale, and O. de Jesús, Neural Network Design (2nd
Edition). Martin Hagan, 2014.

[189] O. A. Mohammed, D. C. Park, F. G. Uler, and C. Ziqiang, “Design optimization


of electromagnetic devices using artificial neural networks,” IEEE Transactions
on Magnetics, vol. 28, no. 5, pp. 2805–2807, 1992.

[190] S. R. H. Hoole, “Artificial neural networks in the solution of inverse


electromagnetic field problems,” IEEE Transactions on Magnetics, vol. 29,
no. 2, 1993.

[191] G. Tsekouras, S. Kiartzis, A. G. Kladas, and J. A. Tegopoulos, “Neural network


approach compared to sensitivity analysis based on finite element technique
for optimization of permanent magnet generators,” IEEE Transactions on
Magnetics, vol. 37, no. 5, pp. 3618–3621, 2001.

[192] E. Snelling, Soft ferrites: properties and applications. Butterworths, 1988.

[193] Y. Wang, D. Ionel, and D. Staton, “Ultrafast Steady-State Multiphysics Model


for PM and Synchronous Reluctance Machines,” Industry Applications, IEEE
Transactions on, vol. 51, no. 5, pp. 3639–3646, 2015.

[194] W. Jiang and T. Jahns, “Coupled electromagnetic-thermal analysis of electric


machines including transient operation based on finite element techniques,” in
Energy Conversion Congress and Exposition (ECCE), 2013 IEEE, 2013, pp.
4356–4363.

[195] M. Popescu and D. Dorrell, “Proximity Losses in the Windings of High Speed
Brushless Permanent Magnet AC Motors With Single Tooth Windings and
Parallel Paths,” Magnetics, IEEE Transactions on, vol. 49, no. 7, pp. 3913–
3916, 2013.

[196] L. Wu, Z. Zhu, D. Staton, M. Popescu, and D. Hawkins, “Analytical Model of


Eddy Current Loss in Windings of Permanent-Magnet Machines Accounting for
Load,” Magnetics, IEEE Transactions on, vol. 48, no. 7, pp. 2138–2151, 2012.

[197] A. Thomas, Z. Zhu, and G. Jewell, “Proximity Loss Study In High Speed Flux-
Switching Permanent Magnet Machine,” Magnetics, IEEE Transactions on,
262

vol. 45, no. 10, pp. 4748–4751, 2009.

[198] P. Reddy, Z. Zhu, S.-H. Han, and T. Jahns, “Strand-level proximity losses in
PM machines designed for high-speed operation,” in Electrical Machines, 2008.
ICEM 2008. 18th International Conference on, 2008, pp. 1–6.

[199] Y. Amara, P. Reghem, and G. Barakat, “Analytical Prediction of Eddy-Current


Loss in Armature Windings of Permanent Magnet Brushless AC Machines,”
Magnetics, IEEE Transactions on, vol. 46, no. 8, pp. 3481–3484, 2010.

[200] L. Wu and Z. Zhu, “Analytical investigation of open-circuit eddy current loss


in windings of PM machines,” in Electrical Machines (ICEM), 2012 XXth
International Conference on, 2012, pp. 2759–2765.

[201] P. Reddy, T. Jahns, and T. Bohn, “Modeling and analysis of proximity losses in
high-speed surface permanent magnet machines with concentrated windings,”
in Energy Conversion Congress and Exposition (ECCE), 2010 IEEE, 2010, pp.
996–1003.

[202] W. Zhang and T. Jahns, “Analytical 2-D slot model for predicting AC losses
in bar-wound machine windings due to armature reaction,” in Transportation
Electrification Conference and Expo (ITEC), 2014 IEEE, 2014, pp. 1–6.

[203] A. Bellara, H. Bali, R. Belfkira, Y. Amara, and G. Barakat, “Analytical


Prediction of Open-Circuit Eddy-Current Loss in Series Double Excitation
Synchronous Machines,” Magnetics, IEEE Transactions on, vol. 47, no. 9, pp.
2261–2268, 2011.

[204] S. Iwasaki, R. Deodhar, Y. Liu, A. Pride, Z. Zhu, and J. Bremner, “Influence of


PWM on the Proximity Loss in Permanent-Magnet Brushless AC Machines,”
Industry Applications, IEEE Transactions on, vol. 45, no. 4, pp. 1359–1367,
2009.

[205] A. Arkadan, R. Vyas, J. Vaidya, and M. Shah, “Effect of toothless stator


design and core and stator conductors eddy current losses in permanent magnet
generators,” Energy Conversion, IEEE Transactions on, vol. 7, no. 1, pp. 231–
237, 1992.

[206] R.-J. Wang and M. Kamper, “Calculation of eddy current loss in axial field
permanent-magnet machine with coreless stator,” Energy Conversion, IEEE
Transactions on, vol. 19, no. 3, pp. 532–538, 2004.

[207] R. Wrobel, J. Goss, A. Mlot, and P. Mellor, “Design Considerations of a


Brushless Open-Slot Radial-Flux PM Hub Motor,” Industry Applications, IEEE
Transactions on, vol. 50, no. 3, pp. 1757–1767, 2014.
263

[208] R. Wrobel, D. Salt, A. Griffo, N. Simpson, and P. Mellor, “Derivation and


Scaling of AC Copper Loss in Thermal Modeling of Electrical Machines,”
Industrial Electronics, IEEE Transactions on, vol. 61, no. 8, pp. 4412–4420,
2014.

[209] P. Mellor, R. Wrobel, D. Salt, and A. Griffo, “Experimental and analytical


determination of proximity losses in a high-speed PM machine,” in Energy
Conversion Congress and Exposition (ECCE), 2013 IEEE, 2013, pp. 3504–3511.

[210] G. Carter, The electromagnetic field in its engineering aspects, ser. Electrical
engineering series. American Elsevier Pub. Co., 1967.

[211] A. Binder, T. Schneider, and M. Klohr, “Fixation of buried and surface-


mounted magnets in high-speed permanent-magnet synchronous machines,”
IEEE Transactions on Industry Applications, vol. 42, no. 4, pp. 1031–1037,
2006.

[212] W. Fei, P. C. K. Luk, and T. S. El-Hasan, “Rotor Integrity Design for a High-
Speed Modular Air-Cored Axial-Flux Permanent-Magnet Generator,” IEEE
Transactions on Industrial Electronics, vol. 58, no. 9, pp. 3848–3858, 2011.

[213] R. Larsonneur, Design and control of active magnetic bearing systems for high
speed rotation. Swiss Federal Institute of Technology, 1990.

[214] M. Barcaro, G. Meneghetti, and N. Bianchi, “Structural Analysis of the Interior


PM Rotor Considering Both Static and Fatigue Loading,” IEEE Transactions
on Industry Applications, vol. 50, no. 1, pp. 253–260, 2014.
264

APPENDIX I

Parametric Open-Slot Stator Structure

The parametrization of the open-slot stator configuration used in Chapter 3 is

illustrated below.

3
13
αwt /2 6 5 4

8 7 10
12
rro dw
11
45o 9 2
αs 1 ds hy

xd
tL
rg
ww
rsi wt/2

rso

The structure can be characterized by defining the following points:


xp1 = rsi
y = 0
p1

xp2 = rso − hy
y = 0
p2
265


xp3 = rslot cos(AngP 3 ), where AngP 3 = αs /2 − arcsin(wt /2/rslot )
y = r sin(Ang ), wherer = r − h
p3 slot P3 slot so y

xp4 = xp5 + xd , where xd = yp5 (xp5 −OOS )−yp10 (xp5 −OOS )
yp10 +xp5 −OOS
y = y − x , whereOO = w /2/ sin(α /2)
p4 p5 d s t s

xp5 = xp6 + ww
y = y
p5 p6

xp6 = dw + rsi − hg /2
y = (x − OO ) tan(α /2) + t
p6 p6 S s L

xp7 = xp6
y = (x − OO ) tan(α /2)
p7 p6 S s

xp8 = rsi cos(αs /2 − arcsin(wt /2/rsi ))
y = R sin(α /2 − arcsin(w /2/r ))
p8 si s t si

xp15 = xp5 + yp5
y = 0
p15

xp10 = xp5
y = (x − OO ) tan(α /2)
p10 p5 S s

xp11 = xp4 + yp8 sin(αs /2)
y = y − y cos(α /2)
p11 p4 p8 s

xp12 = xp13 + yp8 sin(αs /2)
y = y − y cos(α /2)
p12 p13 p8 s

xp13 = xp2 − (xp2 − OOS ) sin(αs /2) sin(αs /2)
y = (x − OO ) sin(α /2) cos(α /2)
p13 p2 S s s
266

Parametric Spoke-Type PM Layout

The parametrization of the spoke-type PM layout used in Chapter 6 is illustrated

below.

1 2
3 4
hpm
αp αpm wbr
wpm
rpm dbr

rro

The structure can be characterized by defining the following points:


xq1 = xq4 − dbr − wpm
y = h /2
q1 pm

xq2 = xq4 − dbr
y = h /2
q2 pm

xq3 = xq4 − dbr
y = w /2
q3 br
 p
xq4 = R2 − (wbr /2)2
ro
y = w /2
q4 br
267

APPENDIX II

Structural FEA for Sensitivity Analysis

The following static structural FEA on the V-type rotor layout was carried out

in Section 7.3 for sensitivity analysis of the stress to the variations of the design

parameters. Values of kdpm , kwpm , kwq , hpm , αpm , and rro are listed in coded form in

the table below for each design.

Design No. hpm kdpm kwpm kwq αpm rro


1 -1 -1 -1 -1 -1 -1
2 -1 -1 -1 -1 1 1
3 -1 -1 -1 1 -1 1
4 -1 -1 -1 1 1 -1
5 -1 -1 1 -1 -1 1
6 -1 -1 1 -1 1 -1
7 -1 -1 1 1 -1 -1
8 -1 -1 1 1 1 1
9 -1 1 -1 -1 -1 1
10 -1 1 -1 -1 1 -1
11 -1 1 -1 1 -1 -1
12 -1 1 -1 1 1 1
13 -1 1 1 -1 -1 -1
14 -1 1 1 -1 1 1
15 -1 1 1 1 -1 1
268

Design No. hpm kdpm kwpm kwq αpm rro


16 -1 1 1 1 1 -1
17 1 -1 -1 -1 -1 1
18 1 -1 -1 -1 1 -1
19 1 -1 -1 1 -1 -1
20 1 -1 -1 1 1 1
21 1 -1 1 -1 -1 -1
22 1 -1 1 -1 1 1
23 1 -1 1 1 -1 1
24 1 -1 1 1 1 -1
25 1 1 -1 -1 -1 -1
26 1 1 -1 -1 1 1
27 1 1 -1 1 -1 1
28 1 1 -1 1 1 -1
29 1 1 1 -1 -1 1
30 1 1 1 -1 1 -1
31 1 1 1 1 -1 -1
32 1 1 1 1 1 1
33 -1 0 0 0 0 0
34 1 0 0 0 0 0
35 0 -1 0 0 0 0
36 0 1 0 0 0 0
37 0 0 -1 0 0 0
38 0 0 1 0 0 0
39 0 0 0 -1 0 0
40 0 0 0 1 0 0
41 0 0 0 0 -1 0
42 0 0 0 0 1 0
43 0 0 0 0 0 -1
44 0 0 0 0 0 1
45 0 0 0 0 0 0
269
270
271
272

APPENDIX III

A Magnetic vector potential (Weber per meter)


ACEEE American Council for an Energy-Efficient Economy
ADVISOR Advanced Vehicle Simulator
AF e Area of iron per pole (square meter)
αpm Magnet pole coverage (degrees)
αs Slot pitch (mechanical radians)
AMC Active Material Cost
ANN Artificial Neural Network
Apm Area of magnet per pole (square meter)
B Magnetic flux density (Tesla)
β Regression coefficients
Bg Peak air-gp flux density (Tesla)
Bpm Average flux density over magnet piece (Tesla)
Br Magnet retentivity (Tesla)
ci Coded design variables
CCD Central Composite Design
CCW Counterclockwise
CE-FEA Computationally Efficient-Finite Element Analysis
CEM Computational Electromagnetics
CMODE Combined Multi-Objective Optimization with Differential Evolution
COP Constrained Optimization Problem
Cr Crossover probability
DE Differential Evolution
δ Skin depth (meter)
DF e Mass density of iron (kilogram per cubic meter)
DOE Design Of Experiments
273

dpm Magnet burial depth (meter)


Dpm Mass density of magnet (kilogram per cubic meter)
Dy Dysprosium
EA Evolutionary Algorithms
EM Electromagnetic
EMF Electromotive Force
EPA Environmental Protection Agency
EPAct Energy Policy and Conservation Act
η2k−1 Skin effect coefficient
EV Electric Vehicles
F Difference scale factor
FC Fan-Cooled
FD Finite Difference
FE Finite Element
FEs Function Evaluations
FEA Finite Element Analysis
FEM Finite Element Method
FSCW Fractional Slot Concentrated Winding
FSM Flux Switching Motor
GA Genetic Algorithm
hg Air-gap height (meter)
hpm Magnet height (meter)
hy Yoke height (meter)
HPC High Performance Computing
HSD Hybrid Synergy Drive
HWFET Highway Fuel Economy Test Driving Schedule
ICH Characteristic current (Ampere)
id D-axis component of armature current phasor (Ampere)
IEC International Electrotechnical Commission
iq Q-axis component of armature current phasor (Ampere)
274

IR Rated current (Ampere)


IPM Interior Permanent Magnet
J Electric current density (Ampere per square meter)
JP M Equivalent electric current density used to represent a permanent
magnet excitation (Ampere per square meter)
kdpm Ratio of magnet burial depth to its maximum possible depth
kh Lamination hysteresis loss coefficient (Watts per Hertz per square
Tesla per kilogram)
ke Lamination eddy-current loss coefficient (Watts per square Hertz
per square Tesla per kilogram)
krac Ratio of ac to dc resistance
ksi Split ratio or the ratio of the stator inner to outer diameter
kσ Rotor leakage coefficient
kwpm Ratio of magnet width to its maximum possible width
kwq Ratio of q-axis web to its maximum possible width
kwt Ratio of tooth-stem width to slot-pitch
kwtt Ratio of tooth-tip-width to slot opening
LA99 Unified Dynamometer Driving Schedule
λ Instantaneous flux-linkage (Weber)
λd D-axis component of flux-linkage (Weber)
λP M Magnet flux-linkage (Weber)
λq Q-axis component of flux-linkage (Weber)
λR Total phase flux-linkage (Weber)
LC Liquid-Cooled
Ld D-axis synchronous inductance (Henry)
Lq Q-axis synchronous inductance (Henry)
m Mass (kilogram)
µ Permeability (Henry per meter)
µr Relative permeability
MC Monte-Carlo
MTPA Maximum Torque Per Ampere
NC Naturally Cooled
275

Nd Neodymium
NdFeB Neodymium Iron Boron
NEDC New European Driving Cycle
NEMA National Electrical Manufacturers Association
Np Population size
NREL National Renewable Energy Laboratory
ν Reluctivity (meter per Henry)
ω Angular frequency (electrical radians per second)
ωm Angular frequency (mechanical radians per second)
ORNL Oak Ridge National Laboratory
p Number of design parameters
P Number of poles
PC Personal Computer
Pac Ac copper losses (Watts)
Pdc Dc copper losses (Watts)
PEV Plug-in Electric Vehicles
P~g Optimization population vector
Pe Eddy-current loss (Watt per kilogram)
PF e Lamination core losses (Watt)
Ph Hysteresis loss (Watt per kilogram)
PM Permanent Magnet
PMSM Permanent Magnet Synchronous Machine
Pw Aggregate weighted loss per unit output power
PWM Pulse Width Modulation
Rac Ac resistance (Ohm)
ρ Resistivity (Ohm meter)
Rdc Dc resistance (Ohm)
rsi Stator inner radius (meter)
rso Stator outer radius (meter)
RSM Response Surface Methodology
SD Steepest Descent
276

SF Slot Fill
σt Tangential stress (Pascal)
SRM Switched Reluctance Motor
SyRM Synchronous Reluctance Motor
Tavg Average torque (Newton meter)
Tem Developed electromagnetic torque profile (Newton meter)
T emp Temperature (degree Celsius)
θ Angle (electrical radians)
θm Angle (mechanical radians)
THD Total Harmonic Distortion
Tr Torque ripple (%)
TRW Torque Ratio per Weight
TSFE Time-stepping FE
~u Trial design members
UDDS Urban Dynamometer Driving Schedule
US06 Supplemental Federal Test Procedure Driving Schedule
vCu Volume of copper (cubic meters)
vR Total induced voltage in the stator winding (Volts)
wpm Width of magnet (meter)
wq Width of q-axis web (meter)
wso Width of slot opening (meter)
wt Tooth width (meter)
wtip Width of tooth tip (meter)
xp Coordinate direction in rectangular Cartesian coordinates
xq Coordinate direction in rectangular Cartesian coordinates
~x Design members
xi,max Upper bound of the design parameters
xi,min Lower bound of the design parameters
~xg Population members
yp Coordinate direction in rectangular Cartesian coordinates
yq Coordinate direction in rectangular Cartesian coordinates

You might also like