Energy Efficiency Improvement of Electric Machines Without RareEarth Magnets

Download as pdf or txt
Download as pdf or txt
You are on page 1of 196

Special Issue Reprint

Energy Efficiency
Improvement of Electric
Machines without
Rare-Earth Magnets

Edited by
Vladimir Prakht, Mohamed N. Ibrahim and Vadim Kazakbaev

www.mdpi.com/journal/energies
Energy Efficiency Improvement of
Electric Machines without
Rare-Earth Magnets
Energy Efficiency Improvement of
Electric Machines without
Rare-Earth Magnets

Editors
Vladimir Prakht
Mohamed N. Ibrahim
Vadim Kazakbaev

MDPI • Basel • Beijing • Wuhan • Barcelona • Belgrade • Manchester • Tokyo • Cluj • Tianjin
Editors
Vladimir Prakht Mohamed N. Ibrahim Vadim Kazakbaev
Ural Federal University Ghent University Ural Federal University
Russia Belgium Russia

Editorial Office
MDPI
St. Alban-Anlage 66
4052 Basel, Switzerland

This is a reprint of articles from the Special Issue published online in the open access journal
Energies (ISSN 1996-1073) (available at: https://www.mdpi.com/journal/energies/special issues/
Energy-Efficiency-Improvement).

For citation purposes, cite each article independently as indicated on the article page online and as
indicated below:

LastName, A.A.; LastName, B.B.; LastName, C.C. Article Title. Journal Name Year, Volume Number,
Page Range.

ISBN 978-3-0365-7656-5 (Hbk)


ISBN 978-3-0365-7657-2 (PDF)

© 2023 by the authors. Articles in this book are Open Access and distributed under the Creative
Commons Attribution (CC BY) license, which allows users to download, copy and build upon
published articles, as long as the author and publisher are properly credited, which ensures maximum
dissemination and a wider impact of our publications.
The book as a whole is distributed by MDPI under the terms and conditions of the Creative Commons
license CC BY-NC-ND.
Contents

About the Editors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Preface to “Energy Efficiency Improvement of Electric Machines without Rare-Earth Magnets” ix

Vladimir Prakht, Mohamed N. Ibrahim and Vadim Kazakbaev


Energy Efficiency Improvement of Electric Machines without Rare-Earth Magnets
Reprinted from: Energies 2023, 16, 3573, doi:10.3390/en16083573 . . . . . . . . . . . . . . . . . . . 1

Vadim Kazakbaev, Vladimir Prakht, Vladimir Dmitrievskii, Safarbek Oshurbekov


and Dmitry Golovanov
Life Cycle Energy Cost Assessment for Pump Units with Various Types of Line-Start Operating
Motors Including Cable Losses
Reprinted from: Energies 2020, 13, 3546, doi:10.3390/en13143546 . . . . . . . . . . . . . . . . . . . 5

Jouda Arfaoui, Hegazy Rezk, Mujahed Al-Dhaifallah, Mohamed N. Ibrahim


and Mami Abdelkader
Simulation-Based Coyote Optimization Algorithm to Determine Gains of PI Controller for
Enhancing the Performance of Solar PV Water-Pumping System
Reprinted from: Energies 2020, 13, 4473, doi:10.3390/en13174473 . . . . . . . . . . . . . . . . . . . 19

Fatma Keskin Arabul, Ibrahim Senol and Yasemin Oner


Performance Analysis of Axial-Flux Induction Motor with Skewed Rotor
Reprinted from: Energies 2020, 13, 4991, doi:10.3390/en13194991 . . . . . . . . . . . . . . . . . . . 37

Bharathi Manne, Malligunta Kiran Kumar and Udochukwu B. Akuru


Design and Performance Assessment of a Small-Scale Ferrite-PM Flux Reversal Wind Generator
Reprinted from: Energies 2020, 13, 5565, doi:10.3390/en13215565 . . . . . . . . . . . . . . . . . . . 53

Hyunwoo Kim, Yeji Park, Seung-Taek Oh, Hyungkwan Jang, Sung-Hong Won,
Yon-Do Chun and Ju Lee
A Study on the Rotor Design of Line Start Synchronous Reluctance Motor for IE4 Efficiency and
Improving Power Factor
Reprinted from: Energies 2020, 13, 5774, doi:10.3390/en13215774 . . . . . . . . . . . . . . . . . . . 79

Ruchao Pupadubsin, Seubsuang Kachapornkul, Prapon Jitkreeyarn, Pakasit Somsiri


and Kanokvate Tungpimolrut
Investigation of Torque Performance and Flux Reversal Reduction of a Three-Phase 12/8
Switched Reluctance Motor Based on Winding Arrangement
Reprinted from: Energies 2022, 15, 284, doi:10.3390/en15010284 . . . . . . . . . . . . . . . . . . . . 95

Seubsuang Kachapornkul, Ruchao Pupadubsin, Pakasit Somsiri, Prapon Jitkreeyarn


and Kanokvate Tungpimolrut
Performance Improvement of a Switched Reluctance Motor and Drive System Designed for an
Electric Motorcycle
Reprinted from: Energies 2022, 15, 694, doi:10.3390/en15030694 . . . . . . . . . . . . . . . . . . . . 113

Himayat Ullah Jan, Faisal Khan, Basharat Ullah, Muhammad Qasim, Ahmad H. Milyani
and Abdullah Ahmed Azhari
Design and Thermal Analysis of Linear Hybrid Excited Flux Switching Machine Using
Ferrite Magnets
Reprinted from: Energies 2022, 15, 5275, doi:10.3390/en15145275 . . . . . . . . . . . . . . . . . . . 131

v
Siddique Akbar, Faisal Khan, Wasiq Ullah, Basharat Ullah, Ahmad H. Milyani
and Abdullah Ahmed Azhari
Performance Analysis and Optimization of a Novel Outer Rotor Field-Excited Flux-Switching
Machine with Combined Semi-Closed and Open Slots Stator
Reprinted from: Energies 2022, 15, 7531, doi:10.3390/en15207531 . . . . . . . . . . . . . . . . . . . 149

Vijina Abhijith, M. J. Hossain, Gang Lei, Premlal Ajikumar Sreelekha,


Tibinmon Pulimoottil Monichan and Sree Venkateswara Rao
Hybrid Switched Reluctance Motors for Electric Vehicle Applications with High Torque
Capability without Permanent Magnet
Reprinted from: Energies 2022, 15, 7931, doi:10.3390/en15217931 . . . . . . . . . . . . . . . . . . . 167

vi
About the Editors
Vladimir Prakht
Vladimir Prakht received a graduate degree in engineering and a Ph.D. degree from the
Department of Electrical Engineering, Ural Federal University, Yekaterinburg, Russia, in 2004
and 2007, respectively. He submitted his Ph.D. dissertation on optimal control and mathematical
modeling induction heating systems in 2006. He is an Associate Professor at the Department of
Electrical Engineering, Ural Federal University. His research interests include the mathematical
modeling and optimal design of energy-efficient electric motors and generators.

Mohamed N. Ibrahim
Mohamed N. Ibrahim (Senior member, IEEE) received an M.Sc. degree in electrical power and
machines engineering from Tanta University, Egypt in 2012, and a Ph.D. degree in electromechanical
engineering from Ghent University, Belgium in 2017. In 2008, he became a Teaching Assistant at
the Electrical Engineering Department, Kafrelshiekh University, Egypt and since 2018, he has been
an Assistant Professor (on leave). Currently, he is a Post-Doctoral Researcher with the Department
of Electromechanical Systems and Metal Engineering, Ghent University, Belgium. He is also an
Affiliate Member of Flanders Make, the strategic research center for the manufacturing industry in
Flanders, Belgium. His major research interests include the design and control of electrical machines
and drives for industrial and sustainable energy applications. Dr. Ibrahim has been a recipient
of the Kafrelshiekh University Award several times for his international scientific publications.
He serves as a reviewer for several journals and conferences, including IEEE Trans. on Industrial
Electronics, Industry Applications, Magnetics, etc. He serves as a Guest Editor for Energies, Mathematics
and Electronics.

Vadim Kazakbaev
Vadim Kazakbaev received a graduate degree in engineering and a Ph.D. degree from the
Department of Electrical Machines, Ural Federal University, Yekaterinburg, Russia, in 2010 and
2017, respectively. He submitted his Ph.D. dissertation on “Development of High-Performance
Synchronous Reluctance Motor” in 2016. He is a Junior Researcher and an Associate Professor with
the Department of Electrical Engineering, Ural Federal University. His research interests include
electrical engineering, the design of electrical machines, and the control of electrical drives.

vii
Preface to “Energy Efficiency Improvement of Electric
Machines without Rare-Earth Magnets”
Electric motors consume about 70% of industrial electricity and about 40%–45% of produced
electricity in the world. This means that using high-efficiency electric motors will reduce the level of
energy consumption and the environmental impact, resulting in reductions in costs and the emissions
of CO2 . In addition, it will significantly reduce the need for new power plants, thus reducing the
invested resources to do so. Electric machines with rare-earth magnets have the highest efficiency
and power density.
However, rare-earth magnets are expensive, and their manufacturing process, as well as
the process of mining rare-earth raw materials, is harmful to the environment. Therefore, the
development of energy-efficient electric machines without rare-earth magnets is of great interest.
Topics of interest for this Special Issue include, but are not limited to:

- The modeling and design of energy-efficient electric machines without rare-earth magnets;
- Synchronous reluctance machines;
- Permanent-magnet-assisted synchronous reluctance machines;
- Switched reluctance machines;
- Hybrid switched reluctance machines;
- DC-excited flux-switching machines;
- PM-excited flux-switching machines;
- DC-excited synchronous machines;
- Direct-on-line and line start motors;
- Induction machines;
- Electric generators without rare-earth magnets for wind turbines;
- Other electric machines without rare-earth magnets;
- Control techniques for electric machines without rare-earth magnets.

Vladimir Prakht, Mohamed N. Ibrahim, and Vadim Kazakbaev


Editors

ix
energies
Editorial
Energy Efficiency Improvement of Electric Machines without
Rare-Earth Magnets
Vladimir Prakht 1 , Mohamed N. Ibrahim 2,3,4, * and Vadim Kazakbaev 1

1 Department of Electrical Engineering and Electric Technology Systems, Ural Federal University,
620002 Ekaterinburg, Russia; [email protected] (V.P.); [email protected] (V.K.)
2 Department of Electromechanical, Systems and Metal Engineering, Ghent University, 9000 Ghent, Belgium
3 Flanders Make @UGent—Core Lab MIRO, 3001 Leuven, Belgium
4 Electrical Engineering Department, Kafrelshiekh University, Kafrelshiekh 33511, Egypt
* Correspondence: [email protected]

Electric motors are one of the largest consumers of electricity and are responsible
for 40–45% of the world’s energy consumption. In some applications, this proportion is
even higher. For example, in industry, electric motors are responsible for about 70% of
electricity consumption. Therefore, reducing the energy consumption of electric motors
is critical both in terms of saving limited fossil fuels and in terms of reducing the load on
national electricity networks. In addition, reducing energy consumption is important in
terms of mitigating the impact of human activities on the environment. It is known that the
magnetic flux generated by a coil with current is proportional to the product of the area of
the pole covered by the coil and the cross-section of the winding wire, that is, the linear
dimensions of the coil to the fourth power. The flux created by a permanent magnet is
proportional to its volume, that is, its linear dimensions to the third degree. For this reason,
electrical machines with high-coercivity rare-earth permanent magnets are potentially the
most compact and energy efficient. At the same time, the cost of rare earth magnets is high,
they are supplied by only a few suppliers, the raw material extraction process for their
manufacture is harmful to the environment and there are problems with disposal. This
gives good motivation for developing electric machines without rare earth magnets [1–10].
This Special Issue aimed to gather new research publications on various topics related
Citation: Prakht, V.; Ibrahim, M.N.; to improving the energy efficiency of electric machines without using rare-earth magnets.
Kazakbaev, V. Energy Efficiency Ten articles have been published that cover various topics. In [1], it is shown that syn-
Improvement of Electric Machines chronous motors without permanent magnets can be an energy-efficient alternative to
without Rare-Earth Magnets. Energies permanent magnet motors, as well as to conventional induction motors in widespread
2023, 16, 3573. https://doi.org/ applications such as fixed-speed electrically driven pump units. The energy-efficiency
10.3390/en16083573 indicators of a direct-on-line synchronous reluctance motor with a power rating of 4 kW
are compared with a motor with rare-earth permanent magnets on the rotor, as well as
Received: 2 March 2023
Revised: 6 April 2023
with an IE3 energy efficiency class induction motor. For this analysis, the performance
Accepted: 18 April 2023
curves of the motors are interpolated based on experimental data. To consider the influence
Published: 20 April 2023
of the power factor of the motors with direct power supply from the grid, losses in the
supply cable of the pumping station were also taken into account. It is shown that despite
the increased losses in the supply cable due to a reduced power factor, the synchronous
reluctance motor provides a significant energy-saving effect compared to the induction
Copyright: © 2023 by the authors. motor due to much higher motor efficiency. The permanent magnet motor provides slightly
Licensee MDPI, Basel, Switzerland. greater energy savings; however, unlike the synchronous reluctance motor, it has a long
This article is an open access article payback period due to its high cost associated with the use of expensive rare earth magnets.
distributed under the terms and The study in [2] is devoted to increasing the energy efficiency of a solar water pump with
conditions of the Creative Commons an electric drive by optimizing the parameters of its PI controllers. The solar pump is a
Attribution (CC BY) license (https:// self-powering system that does not require power from the utility network. The considered
creativecommons.org/licenses/by/
electric drive uses an induction motor that does not have expensive permanent magnets.
4.0/).

Energies 2023, 16, 3573. https://doi.org/10.3390/en16083573 1 https://www.mdpi.com/journal/energies


Energies 2023, 16, 3573

The control system uses DC link voltage and motor speed PI controllers. It is shown
that, when tuning the PI controllers, the coyote optimization algorithm is more efficient
in terms of computational costs and the final result than the widely used Ziegler–Nichols
and trial error tuning algorithms. Paper [3] theoretically discusses the design of an axial
flux induction motor without permanent magnets, the magnetic core of which is made
of Soft Magnetic Composites (SMC), which simplifies its production. It is shown that
such a motor can correspond to the IE3 energy efficiency class. It is also highlighted that
rotor skew can provide low torque ripple for such a motor. In [4], the choice of the best
flux reversal machine (FRM) configuration for a small wind generator with inexpensive
ferrite magnets is discussed. A three-phase FRM design with six stator winding poles and
eight rotor teeth with significantly reduced cogging torque and torque ripple is selected
for this application based on comparative finite element analysis (FEA). In [5], manual
optimization of the parameters of the rotor geometry of a synchronous reluctance motor
without magnets with a direct start from the grid is carried out. As a result, a significant
increase in efficiency and power factor is achieved. The results of testing the motor at the
rated load are provided, verifying the theoretical results obtained. In [6], two winding
layouts of a three-phase switched reluctance motor (SRM) with twelve stator teeth and
eight rotor teeth are compared in terms of power loss and efficiency. It is shown that
the winding with a short pitch provides a lower copper of loss than the winding with a
full pitch. Comprehensive comparative experiments were carried out to verify the results
obtained using FEA. In [7], various aspects of the development and manufacture of an
electric motorcycle electric drive based on an SRM are discussed. The choice of SRM
configuration for this application is discussed. It is shown that a three-phase SRM with
six stator teeth and four rotor teeth shows the best performance for this application. The
dynamic simulation of SRM as part of the drive of the electric vehicle is carried out. A
method for determining the position of the rotor and its experimental implementation are
discussed. The test results of the SRM under consideration on the test bench and as part of
an electric motorcycle are also provided. In [8], a novel design of a linear flux switching
motor (FSM) is proposed employing inexpensive ferrite magnets with a modular passive
stator for railway transport. Optimization of the FSM parameters using a genetic algorithm
was carried out. The results of the electromagnetic and thermal analysis are presented.
The proposed design of the linear FSM has a high specific thrust and high efficiency and
a wide range of velocity control. In [9], a novel FSM design without permanent magnets
with an external rotor is proposed. The outer rotor is passive and has no windings. Due to
the selected shape of the stator teeth, the proposed FSM has a reduced leakage flux, high
power factor and low torque ripple. The FSM design is optimized using a genetic algo-
rithm. The results of electromagnetic and thermal finite element analysis are presented. A
theoretical comparison of the performance of the optimized FSM with several conventional
wound rotor synchronous machines and FSM configurations is provided, which shows the
superiority of the proposed solution in terms of efficiency and specific torque. In [10], for
electric vehicles, a novel design of an SRM with twelve stator teeth and eight rotor teeth
without permanent magnets is proposed. Unlike conventional SRMs, the proposed motor
uses a hybrid winding excitation algorithm that provides increased specific torque. The
results of FEA are confirmed by comprehensive experimental verification carried out at
various speeds and torques of the motor. It is shown that the specific torque and efficiency
of the proposed hybrid SRM are significantly higher than that of the conventional one.

Author Contributions: Conceptualization, M.N.I., V.P. and V.K.; methodology, M.N.I.; software,
V.K.; validation, V.K., M.N.I. and V.P.; formal analysis, V.K.; investigation, V.K.; resources, V.P.; data
curation, V.K.; writing—original draft preparation, M.N.I.; writing—review and editing, M.N.I., V.K.
and V.P.; visualization, V.K.; supervision, V.P. All authors have read and agreed to the published
version of the manuscript.
Acknowledgments: The editors thank the authors for submitting their studies to the present Special
Issue which led to a successful completion. We deeply acknowledge the Energies Reviewers for

2
Energies 2023, 16, 3573

enhancing the quality and impact of all submitted papers. Finally, we sincerely thank the editorial staff
of Energies for their stunning support during the development and publication of this Special Issue.
Conflicts of Interest: The author declares no conflict of interest.

References
1. Kazakbaev, V.; Prakht, V.; Dmitrievskii, V.; Oshurbekov, S.; Golovanov, D. Life Cycle Energy Cost Assessment for Pump Units
with Various Types of Line-Start Operating Motors Including Cable Losses. Energies 2020, 13, 3546. [CrossRef]
2. Arfaoui, J.; Rezk, H.; Al-Dhaifallah, M.; Ibrahim, M.N.; Abdelkader, M. Simulation-Based Coyote Optimization Algorithm to
Determine Gains of PI Controller for Enhancing the Performance of Solar PV Water-Pumping System. Energies 2020, 13, 4473.
[CrossRef]
3. Arabul, F.K.; Senol, I.; Oner, Y. Performance Analysis of Axial-Flux Induction Motor with Skewed Rotor. Energies 2020, 13, 4991.
[CrossRef]
4. Manne, B.; Kumar, M.K.; Akuru, U.B. Design and Performance Assessment of a Small-Scale Ferrite-PM Flux Reversal Wind
Generator. Energies 2020, 13, 5565. [CrossRef]
5. Kim, H.; Park, Y.; Oh, S.T.; Jang, H.; Won, S.H.; Chun, Y.D.; Lee, J. A Study on the Rotor Design of Line Start Synchronous
Reluctance Motor for IE4 Efficiency and Improving Power Factor. Energies 2020, 13, 5774. [CrossRef]
6. Pupadubsin, R.; Kachapornkul, S.; Jitkreeyarn, P.; Somsiri, P.; Tungpimolrut, K. Investigation of Torque Performance and Flux
Reversal Reduction of a Three-Phase 12/8 Switched Reluctance Motor Based on Winding Arrangement. Energies 2022, 15, 284.
[CrossRef]
7. Kachapornkul, S.; Pupadubsin, R.; Somsiri, P.; Jitkreeyarn, P.; Tungpimolrut, K. Performance Improvement of a Switched
Reluctance Motor and Drive System Designed for an Electric Motorcycle. Energies 2022, 15, 694. [CrossRef]
8. Jan, H.U.; Khan, F.; Ullah, B.; Qasim, M.; Milyani, A.H.; Azhari, A.A. Design and Thermal Analysis of Linear Hybrid Excited Flux
Switching Machine Using Ferrite Magnets. Energies 2022, 15, 5275. [CrossRef]
9. Akbar, S.; Khan, F.; Ullah, W.; Ullah, B.; Milyani, A.H.; Azhari, A.A. Performance Analysis and Optimization of a Novel
Outer Rotor Field-Excited Flux-Switching Machine with Combined Semi-Closed and Open Slots Stator. Energies 2022, 15, 7531.
[CrossRef]
10. Abhijith, V.; Hossain, M.J.; Lei, G.; Sreelekha, P.A.; Monichan, T.P.; Rao, S.V. Hybrid Switched Reluctance Motors for Electric
Vehicle Applications with High Torque Capability without Permanent Magnet. Energies 2022, 15, 7931. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

3
energies
Article
Life Cycle Energy Cost Assessment for Pump Units
with Various Types of Line-Start Operating Motors
Including Cable Losses
Vadim Kazakbaev 1 , Vladimir Prakht 1, *, Vladimir Dmitrievskii 1 , Safarbek Oshurbekov 1 and
Dmitry Golovanov 2
1 Department of Electrical Engineering and Electric Technology Systems, Ural Federal University,
Ekaterinburg 620002, Russia; [email protected] (V.K.); [email protected] (V.D.);
[email protected] (S.O.)
2 Department of Electrical Engineering, University of Nottingham, Nottingham NG7 2RD, UK;
[email protected]
* Correspondence: [email protected]

Received: 11 June 2020; Accepted: 6 July 2020; Published: 9 July 2020

Abstract: The paper presents a comparative analysis of life-cycle energy consumption for three
different types of 4 kW line-start motors used in a pump unit with throttling: the most widely used
induction motor with IE3 efficiency class, line start permanent magnet synchronous motor with IE4
efficiency class and line start synchronous reluctance motor with IE4 efficiency class. The operating
cycle for pump units with constant flow is considered for the above-mentioned types of motors taking
into account not only the losses in the pump and motor, but also in the power supply cable. It is shown
that the line start synchronous reluctance motor without magnets has the highest efficiency over the
entire considered loading range. However, its power factor is lower than that of the synchronous
motor with magnets and therefore it has more significant losses in power supply cable. Despite
this disadvantage, the line-start reluctance motor is a good alternative to widespread induction
motor since it allows saving of approximately 4000 euro more than the latter during the 20 years life
cycle. It also provides similar savings in comparison to the permanent magnet synchronous motor,
but unlike it, it does not have costly rare-earth materials in the rotor.

Keywords: centrifugal pump; energy efficiency; induction motor; line-start synchronous motor;
synchronous reluctance motor; throttling control

1. Introduction
Nowadays, most of the drives with the direct start from the mains power supply utilize induction
motors (IM) with aluminium squirrel cage. Typically, IM with casting squirrel cage has a relatively low
class of efficiency IE3 [1,2].
Although it is currently acceptable to use the motors with the efficiency class IE3 for industrial
applications, according to the EU plans, the efficiency requirements will be increased in the future.
For instance, starting from 1st of July 2023, the motors in the EU with the output power greater than
75 kW should comply with the IE4 class of efficiency [3]. Future plans imply the expansion of IE4
requirements to the motors with lower output power (<75 kW) and moving towards the IE5 class of
efficiency for high output power motors [4]. In addition, even at the present time, the use of motors of
IE4 and IE5 classes can be feasible, due to constantly increasing cost of energy resources and the need
to reduce environmental impact [5].
Therefore, further efficiency improvement of IM is necessary in order to comply with upcoming
changes in energy efficiency standards. One of the possible ways is to replace the aluminium squirrel

Energies 2020, 13, 3546; doi:10.3390/en13143546 www.mdpi.com/journal/energies

5
Energies 2020, 13, 3546

cage with a copper one. However, the latter is more expensive in comparison to the aluminium squirrel
cage [2]. For that reason, the IMs with copper squirrel cage are not so widespread for the drives with
direct start from the mains power supply.
Recently, the major industrial companies [6,7] have developed synchronous permanent magnet
motors with the line start from the mains power supply—so-called line start permanent magnet
synchronous motors (LS-PMSM), which comply with the IE4 class efficiency. Such synchronous motors
have a squirrel cage type winding on the rotor. This squirrel cage allows starting the motor directly
from the mains power supply in asynchronous mode. The rotor accelerates until its frequency gets
close to the synchronous frequency. In synchronous mode, the squirrel cage dampens rotor vibrations
during sudden changes in load.
There is a number of publications dedicated to the development of LS-PMSM [8–13]. However,
LS-PMSM cannot compete with IM contribution to industrial use due to the high cost of the former.
Another problem in the manufacturing of LS-PMSM is its technological dependence on rare earth
materials suppliers from China. About 95% of the total amount of rare earth raw materials worldwide
comes from China [14]. Since this situation is unlikely to change in the foreseeable future, there is
always a threat of an unstable change in the price of raw materials for the production of permanent
magnets. The already high prices for materials for the production of rare earth magnets can change
several times within a few years [14,15].
In addition, the rare earth mineral processing technology is associated with significant
environmental damage. It is claimed that the processing of each ton of rare-earth element concentrate
results in generation of 1–1.4 t of radioactive waste. Only a small part of the resulting slush contains
rare-earth elements and is subsequently extracted for refinement [16].
In view of the above-noted disadvantages of the motors with rare earth magnets, an alternative
to the LS-PMSM must be sought. Currently, synchronous reluctance motors (SynRM), which have
no magnets and no short-circuited startup winding in the rotor, and are powered by a variable
frequency drive (VFD) are the good alternatives to IM [17–21]. They meet the criteria of IE4 energy
efficiency [22,23] and IE5 [19,24,25], according to the standard IEC 60034-30-2 «Rotating electrical
machines – Part 30-2: Efficiency classes of variable speed AC motors (IE-code)». These kinds of SynRMs
were recently launched for production by large industrial companies. [22,26,27].
There are also IE4 energy efficiency class synchronous reluctance motors that are powered directly
from the mains power supply, the rotor of which is manufactured with a short-circuited startup
winding (line start synchronous reluctance motor, LS-SynRM) [28–31]. These kinds of LS-SynRMs have
a higher energy efficiency class than IM as well as approximately the same manufacturing cost [32].
Wide use of energy-efficient LS-SynRMs can lead to a decrease in energy consumption and
energy intensity of GDP, as well as lower Greenhouse Gas emissions during electricity generation.
The use of energy-efficient LS-SynRM instead of IM will help to achieve the goals stated in the energy
and environmental strategies of the European Union (European Green Deal [5]), USA (State Energy
Program [33]), Switzerland (supporting Paris Climate Agreement [34]), China (supporting Paris Climate
Agreement [35]), Japan (Net Zero Energy Building [36]), South Korea (supporting Paris Climate
Agreement [37]) and other countries.
Pump systems consume almost 22% of all electric energy generated throughout the world [38].
Most of the pump drives are powered directly from the mains power supply [39,40]. This suggests the
high energy-saving potential of LS-SynRM in pump applications.
A large number of works are dedicated to the comparison of the energy consumption of pumping
systems with various types of motors (induction motors, synchronous motors with rare earth permanent
magnets, synchronous reluctance motors without magnets). However, all these works are dedicated to
pumping systems with motor frequency control using variable frequency drive (VFD) [17,21,41–43].
Energy-saving effect of using various types of motors in pumping systems powered directly from
the mains power supply are considered much less frequently. Thus, in [40] the comparison of the
energy consumption for LS-PMSM and IM with the direct feed from the mains power supply of classes

6
Energies 2020, 13, 3546

IE3 and IE4 in a centrifugal pump with a throttling control is considered. In that work, it was shown
that when choosing the motor, it is necessary to take into account not only the energy efficiency class,
but also the efficiency under reduced loads.
However, the estimation of the energy-saving effect of LS-SynRM use in pumping systems,
and their comparison to IM and LS-PMSM, including the energy losses in power supply cable, has not
yet been considered in the literature to the best of our knowledge. The energy losses in the power
supply cable depend on the power factor. The power factor should be taken into account due to the
increased reactive component of the current, which contributes significantly to the increase of the total
current. In [44–46], it is noted that LS-SynRM has a small power factor, which can lead to losses in the
power supply cable. Since the motors studied in this article have different values of the power factor,
not only their efficiency but also their influence of the power factor on losses in the power supply cable
is taken into account when comparing their energy consumption.
In this paper, a comparative assessment of the energy consumption for a low-power pump drive
(4 kW) with various types of electric motors is carried out. The following electric motor types are
considered: LS-SynRM with energy efficiency class IE4 (Figure 1a), LS-PMSM with energy efficiency
class IE4 (Figure 1b) and IM with energy efficiency class IE3 (Figure 1c). All three considered motors
have a similar design of the stator. However, their rotor design is different. All the motors have a
starter cage on the rotor for the asynchronous startup. However, LS-SynRM has also the magnetic
anisotropy of the rotor structure formed by magnetic barriers. Therefore, it enters into synchronous
operation after starting. LS-PMSM also goes into synchronous operation due to the synchronizing
torque produced by the permanent magnets installed in the rotor. The synchronous motors (LS-SynRM
and LS-PMSM) usually have a better efficiency compared to IM due to the reduced rotor losses. Various
loading modes of the duty cycle of the pump drive with unregulated rotation speed are taken into
account. The costs of electric energy consumption during the pump life-cycle is chosen as the main
criterion for comparing the motors.

(a) (b) (c)

Figure 1. Schematic representation of motor geometry: (a) line start synchronous reluctance motors
(LS-SynRM); (b) line start permanent magnet synchronous motors (LS-PMSM); (c) induction motors (IM).

It must be highlighted that the analysis presented in this paper is not based on theoretical
calculations of motor performances using a modelling software. Instead, an approach to the analysis
of energy consumption based on data from manufacturers' datasheets and experimental data is
used in the work. For IM and LS-PMSM, the data from manufacturers datasheets are used [47,48].
Since LS-SynRMs of high energy efficiency classes are not yet mass-produced, the experimental data for
LS-SynRM are taken from the article [28]. More specifically, the data on the efficiency and power factor
of the motors are used in the present analysis. For data processing, polynomial interpolation is applied
to the entire load range under consideration. Interpolation of experimental data and other calculations
were performed in the MATLAB program and are described in Formulas (1)–(9). The proposed
calculation method can be used for choosing the type of electric motor in pumping systems.

2. Evaluating Pump Energy Consumption


The mathematical model of pump unit for calculating the energy consumption of a pump drive
with various types of the motors is presented below. The schematic representation of the drive of
a pump unit with one electric motor, powered directly from the mains power supply, is shown in

7
Energies 2020, 13, 3546

Figure 2. The electric motor is fed directly from the mains and is coupled to a centrifugal pump without
intermediate mechanical gears. The required motor electrical power Pmotor el depends on the flow
Q [39]:
Pmotor el = ηmotor ·Pmech ; (1)

Pmech = ηpump ·Phydr = ηpump ·ρ·g·H·Q = f(Q) (2)

where H—water pressure, is defined from the H-Q characteristic of the pump from the catalogue;
g—acceleration of gravity; ρ—density of a liquid; Phydr —hydraulic power of pump; Pmech —mechanical
power of pump, defined from characteristic Pmech = f (Q) from catalogue; ηpump —pump efficiency;
ηmotor —motor efficiency.

Figure 2. Diagram of a single pump unit for fixed-speed operation. In this diagram: n, T are the
rotational speed and load torque of the motor; pmotor is the motor loss; ppump is the pump loss.

In addition to losses in the motor, losses in the power supply cable are also taken into account.
The power supply cable electrical loss pcable depends on the active phase resistance of the cable and the
value of the motor current [49]:
pcable = 3·Rcable ·Imotor 2 , (3)

where Rcable —cable phase resistance; Imotor —motor current.


The motor current is calculated according to formula:

Imotor = Pmech /( 3·Vmotor ·cosϕ·ηmotor ), (4)

where Vmotor = 400 V; cosϕ and ηmotor —power factor and motor efficiency, according to data from
the catalogue.
Considering losses in cable, the electric power P1 consumed from the mains by the pump unit is
calculated as:
P1 = ηmotor ·Pmech + pcable . (5)

To calculate the energy consumption of a pump drive, the pump performance characteristics from
the manufacturer's datasheets were used [50]. In order to compare the energy consumption of the
electric motors as a part of the pump unit when controlling flow by a throttling valve, a centrifugal
pump B-NM4 65/25B/B (manufactured by Calpeda S.p.A., Montorso Vicentino, Vicenza, Italy) with the
rated power Prate = 4 kW and with rated rotational speed n = 1450 rpm was used [50]. Pump data are
shown in Table 1. QBEP denotes the flow at the best efficient point (BEP), and HBEP denotes the pump
head at BEP.

Table 1. Published characteristics of the pump from manufacturer.

Rated Mechanical Rate Rotational Pump Efficiency


Type QBEP , m3 /h HBEP , m
Power, W Speed, rpm in BEP, %
B-NM4
4000 1450 60 15.4 75.5
65/25B/B

8
Energies 2020, 13, 3546

The calculation was performed for three different four-pole electric motors with the rated power of
4 kW, namely LS-SynRM (a test prototype [28]), LS-PMSM (manufacturer WEG [47]) and IE3 efficiency
class induction motor (IM, manufacturer WEG [48]). For the serially produced LS-PMSM and IM,
datasheets on their efficiency are used.
There are still no commercially available high-performance LS-SynRMs on the market to the best
of our knowledge. ABB Group corporation has announced the launch of IE4 class LS-SynRM [51],
however, at the moment of writing this manuscript, these motors are still not available on the market.
Therefore, to perform the calculation of LS-SynRM power consumption, the data of the experimental
sample described in [28] were used.
Efficiency data for the motors are shown in Tables 2 and 3 and in Figure 3. Current data of the
motors are shown in Figure 4. Power factor data of the motors are shown in Figure 5. The motor
current can be found based on the efficiency and power factor data using Equation (4).

Table 2. Motor characteristics.

Rated Mechanical Frame


Type of Motor Poles Frame Size Weight, kg Rated Voltage, V
Power, W Material
LS-SynRM 4000 4 IEC 112 No data No data 400
LS-PMSM 4000 4 IEC 112 Cast iron 49 400
IM 4000 4 IEC 112 Cast iron 42 400

Table 3. Motor characteristics.

Motor Efficiency, % Motor Power Factor


Type of Motor
50% Load 75% Load 100% Load 50% Load 75% Load 100% Load
LS-SynRM 91.6 92.4 91.9 0.607 0.713 0.755
LS-PMSM 89.0 91.0 91.7 0.68 0.81 0.88
IM 88.7 89.1 88.8 0.6 0.72 0.78

Figure 3. Motor efficiency.

Figure 4. Motor current.

9
Energies 2020, 13, 3546

Figure 5. Motor power factor.

Note that all three considered motors have approximately the same size and are located in standard
casings IEC 112 with mounting dimensions in accordance with IEC standard 60072-1-1991. All data for
LS-PMSM and IM are taken from the datasheets of the manufacturer [47,48]. Data for LS-SynRM are
taken from the article [28].
In Figures 3–5, the data on motor efficiency, current and power factor are presented. It can be seen
that the LS-SynRM has the highest efficiency values in various modes (Figure 3). The rated efficiency
of the LS-SynRM is 91.9%. Also, LS-PMSM is more efficient than the traditional IM. According to the
data in Table 3, the estimated efficiency of the IM is only 88.8%, while the efficiency of LS-PMSM is
significantly higher at 91.7%. Figure 3 also demonstrates that the LS-PMSM efficiency is higher than
that of the IM over the entire considered loading range.
The efficiency of the LS-PMSM in the nominal mode is the same as that of the LS-SynRM. However,
in partial load modes, the LS-PMSM's efficiency decreases much faster than the one of LS-SynRM.
LS-PMSM has the highest power factor. The power factors of IM and LS-SynRM have approximately
the same values (0.78 and 0.755, respectively) which are significantly lower than the one for PMSM
(0.88). The rate of the power factor decrease with a decrease in load is approximately the same for
all studied motors: when the load was decreased from 100% to 50%, the power factor decreased by
approximately 0.15.

3. Pump Operating Cycle


The operation of the pump unit was considered in the loading points wherein the water flow over
the duty cycle changes in accordance with the typical characteristic of fixed-speed pump applications,
as shown in reference [52]. In reality, however, even in the case of fixed-speed pumps, the flowrate is
very seldom constant. For example, even in a simple pump system when a pump is used to move liquid
from one reservoir to another, the duty points vary due to the level of the reservoirs, which means
the pump does not operate all the time at the best efficiency. A typical duty cycle of a pump with an
approximately constant flowrate described in [52] is characterized by three discrete modes (Figure 6).

Figure 6. Flow-time profile. The numbers in the rectangles indicate the share of the operation time of a
loading point.

10
Energies 2020, 13, 3546

An electric motor is connected directly to the mains power supply. Therefore, the flow Q of the
pump is adjusted using a throttle. In this case, the water pressure changes in accordance with the Q-H
curve of the pump, and the operating point is the intersection point of the pump characteristic (red line
in Figure 7a) and the hydraulic system characteristic (blue, green and emerald lines in Figure 7a).
Figure 7a shows the interpolation results of the Q-H characteristic of the selected pump according to
manufacturer data [50].

(a) (b)

(c)
Figure 7. Pump performances: (a) Q-H curve; (b) pump mechanical power; (c) pump efficiency.

The mechanical (input) power curve of the pump as a function of the flow is reported by the
pump manufacturer (Figure 7b). The pump power was determined from the curve in three operation
modes (75%, 100% and 110% of the pump flow). The flow corresponding to 100% of the maximum
flow in the pump operating cycle was determined based on the pump efficiency curve (Figure 7c) [50],
corresponding to maximum efficiency (best efficiency point: Q = 60m3 /h, ηpump = 0.755).
The efficiencies of the electric motors (Table 2) at the four operation modes of the pump unit were
determined using polynomial interpolation of their specification data.
The obtained efficiency values are provided in Table 4, which also shows the following values
for each operating mode: the flow, the pump head, and the mechanical output power of the electric
motors as percentages of the rated output.

Table 4. Characteristics of pump duty cycle.

ηpump , % ηmotor , %
Q, % Q, m3 /h H, m Pmech , W T, N·m
LS-SynRM LS-PMSM IM
110 66 14.4 3453 22.0 75 0.924 0.915 0.89
100 60 15.4 3335 21.2 75.5 0.925 0.914 0.891
75 45 17.25 2962 18.9 71.4 0.925 0.909 0.891

11
Energies 2020, 13, 3546

4. Cable Losses Depending on Motor Power Factor


An important energy parameter for a motor with a feed from the mains power supply is not only
the efficiency, but also the power factor, because the reactive current fed to a motor flows not only
through its winding, but also through a network of elements from which a motor receives power [53]
causing additional losses. Since the considered motors have different power factors and total currents
(see Figure 4 and Table 5), it is also necessary to evaluate the influence of this factor on the cost of
electricity for the consumer.

Table 5. Motor current at three points of the duty cycle for various pump modes.

Imotor , A
Q, % Pmech , W
LS-SynRM LS-PMSM IM
110 3453 7.33 6.42 7.42
100 3335 7.12 6.27 7.23
75 2962 6.51 5.84 6.69

Let us have a look at the simplest case when a pump is connected directly to a three-phase 400 V
network. In this case, if we want to consider cable losses, we need to take into account the magnitude
of the motor current too. For industry, the typical cable length for connecting low-voltage power
equipment is about 100 m [54]. In low-power electrical systems with a current load of up to 15A, cables
with a cross section of 1.5 mm2 are typically used [55]. The specific resistance of one phase of a copper
cable with these parameters is approximately ρcable = 12.6 Ohm/km. When performing calculations
for the stranded cables of a small cross section, the reactance is usually neglected. We assume that
the motor is powered via a typical cable of 100 m length (lcable = 100 m). The phase resistance of such
a cable will be Rcable = lcable ·ρcable = 0.1·12.6 = 1.26 Ohm. Losses in the cable are calculated from (3).
The results of interpolation of motor current values for various studied pump modes are presented in
Table 5.
Table 6 presents the results of the calculation of losses in the cable according to Formula (3), as well
as the values of the losses in the motor, according to the data in Table 4.

Table 6. Losses in motor and cable at three points of the duty cycle for various pump modes.

pmotor , W pcable , W
Q, % Pmech , W
LS-SynRM LS-PMSM IM LS-SynRM LS-PMSM IM
110 3453 283.6 320.8 426.8 203.0 155.6 208.0
100 3335 272.4 313.8 408.0 191.8 148.6 197.7
75 2962 241.9 296.6 362.4 160.0 128.8 169.4

In accordance with the assumptions made, the estimated losses in cable are comparable with the
ones in the motor (Table 6). At the same time, the cable losses for LS-PMSM are lower by approximately
25% than for LS-SynRM and IM. The main reason of the reduced cable losses for LS-PMSM is its
reduced total current due to its higher power factor, according to Equations (3) and (4). These results
confirm the importance of the power factor increase for reduction of the energy consumption of the
line-start motors.

5. Electric Energy Consumption Results and Discussion


Using the results obtained in the previous sections, we compared the energy consumption of the
pump drive with various motors: LS-SynRM, LS-PMSM and IM motor. The daily energy consumption

12
Energies 2020, 13, 3546

for each electric motor over a full duty cycle of the pump unit in accordance with the corresponding
load profile (Figure 6) is determined as:


3
Eday = tΣ · (P1 i · ti /tΣ ). (6)
i=1

where i = 1 . . . 3 is the number of a loading point; P1 i is the eclectic power P1 in a loading point; ti /tΣ is
the share of the operation time of a loading point.
For the year-round operation of the pump unit, the annual energy consumption was calculated
as follows:
Eyear = Ed ·365. (7)

The cost of electricity consumed (in Euro), considering the applied grid tariffs GT = 0.2036 €/kW·h
for non-household consumers [56] for Germany in the second half of 2019, was calculated as follows:

Cyear = Ey ·GT. (8)

The whole life cycle of pump units is usually about 20 years [57,58]. The energy cost for the
lifespan of n = 20 years is assessed without taking into account the maintenance costs and the initial
cost of the motors since the market cost of the motors depends on many factors and this was beyond
of the topic of the present paper. Furthermore, the pump lifetime expenses often consist mostly of
the energy cost (>50–60%) [57,58]. The net present value (NPV) of the life-cycle cost was calculated
as follows:
 n  
CLCC = C year /[1 + ( y − p)]i , (9)
i=1

where y— interest rate (y = 0.04); p—expected annual inflation (p = 0.02); n—lifetime of the pump unit
(n = 20 years) [58].
The results of calculations based on (1)–(9) excluding and including cable losses are presented in
Tables 7 and 8, respectively. Figure 8 presents calculated values of life cycle energy cost CLCC for both
cases. The energy savings of LS-SynRM and LS-PMSM is assessed in relation to the induction motor.

Table 7. Calculation of energy consumption excluding cable losses.

ti /tΣ , % LS-SynRM LS-PMSM IM


25 3736.7 3773.9 3879.9
P1 , W
50 3607.3 3648.8 3742.9
25 3204.5 3259.2 3325.0
Eday , kW·hour 84.9 86.0 88.1
Eyear , kW·hour 31,001 31,384 32,173
Annual energy savings, kW·hour 1171 789 –
Annual energy savings, % 3.65 2.45 –
Cost savings, € (per year) 238.5 160.6 –
Cost savings, € (per 20 years) 3900 2626 –
Life cycle energy cost CLCC , € (20 years) 103,208 104,482 107,108

13
Energies 2020, 13, 3546

Table 8. Calculation of energy consumption including cable losses.

ti /tΣ , % LS-SynRM LS-PMSM IM


25 3939.7 3929.5 4087.9
P1 , W
50 3799.1 3797.3 3940.7
25 3364.5 3387.9 3494.4
Eday , kW·hour 89.4 89.4 92.8
Eyear , kW·hour 32,636 32,658 33,865
Annual energy savings, kW·hour 1228.9 1207.6 –
Annual energy savings, % 3.6 3.6 –
Cost savings, € (per year) 250.2 245.9 –
Cost savings, € (per 20 years) 4181 4020 –
Life cycle energy cost CLCC , € (20 years) 108,651 108,722 112,742

(a) (b)

Figure 8. Life cycle energy cost (20 years) using various motor types: (a) without considering cable
loss; (b) considering cable loss.

From Tables 7 and 8 and Figure 8 we can see that without taking into account the cable losses,
LS-SynRM provides the lowest energy consumption (Eday = 84.9 and Eyear = 31,001 kW·hour) which is
3.65% more annual energy saving than for IM. For the LS-PMSM and IM the daily energy consumption
values are 86.0 and 88.1 kW hour, respectively, and their annual energy consumption is 31,384 and
32,173 kW hour, respectively. The annual cost saving for LS-PMSM is 2.45% higher than for IM.
When the cable losses are taken into account the energy savings in relation to IM of both LS-SynRM
and LS-PMSM raise to 3.6%.
The life cycle energy cost for 20 years of use is estimated in euros and is 108,651, 108,722 and
112,742 € including cable losses for LS-SynRM, LS-PMSM and IM, respectively. It can be seen that
for LS-SynRM, the life cycle savings in energy costs against the IM excluding and including cable
losses are relatively similar 3900 € and 4181 €, respectively. In case of using LS-PMSM, the savings
increase almost double from 2626 € to 4020 € when the cable losses were taken into account. Thus,
before making a choice of particular motor type and calculating the payback period, it is necessary to
take into account not only the efficiency of the motors, but also the power factor and its effect on cable
losses, as this can significantly affect the results of the feasibility study.
However, as we have already mentioned above, LS-PMSMs have a higher cost compared to IM
and LS-SynRM due to the use of expensive rare-earth magnets in its design. In addition, the rare-earth
elements processing from raw ore is associated with significant environmental damage [16]. Therefore,
the LS-SynRM, which does not have permanent magnets and has a comparable cost to IM, is the most
attractive pumping system in terms of energy saving even taking into account cable losses.

14
Energies 2020, 13, 3546

6. Conclusions
The present paper provides a comparative analysis of the energy consumption of electric motors
of various types (LS-SynRM, LS-PMSM and IM) used as part of a 4 kW fixed-speed pump unit with
throttle control. The comparison takes into account not only the efficiency of the motors at different
pump loads, but also the effect of the motor power factor on cable losses in the power supply line.
It is shown that the life cycle energy cost savings for LS-SynRM compared to IM was 3900 €
excluding cable losses. For LS-PMSM this number was 2626 €. However, when taking into account
the cable losses, affected by the power factor and the total current of the motor, the savings for both
LS-SynRM and LS-PMSM are almost the same: 4181 € and 4020 €, respectively.
Therefore, when selecting the motor type and calculating the payback period, it is necessary to
take into account not only the efficiency of the motors, but also the motor power factor and its impact
on the cable losses, as this can significantly affect the results of the feasibility study.
Thus, the use of LS-SynRM in the considered pump application provides lower cost of motor
manufacturing and environmental friendliness, compared to LS-PMSM, and energy savings of more
than 4000 € during the life cycle, compared to IM.
Based on this study, LS-SynRM can be suggested as the best alternative for the considered pump
application since having the lowest energy consumption. In addition, LS-SynRM has the manufacturing
cost comparable to IMs, while the cost of LS-PMSM is significantly higher, due to the presence of
permanent magnets, the extraction of which is harmful to the environment.
The proposed method can be applied in the analysis of the energy consumption of other motor
types used in pump stations, for example, line-start permanent magnet assisted synchronous reluctance
motors (LS-PMaSynRM).

Author Contributions: Conceptual approach, V.K. and V.P.; data curation, D.G., S.O., and V.K.; software, S.O. and
V.K.; calculations and modelling, S.O., V.D., V.K. and V.P.; writing of original draft, D.G., S.O., V.D., V.K. and V.P.;
visualization, D.G., and V.K.; review and editing, D.G., S.O., V.D., V.K. and V.P. All authors have read and agreed
to the published version of the manuscript.
Funding: This research received no external funding.
Acknowledgments: The authors thank the editors and reviewers for careful reading and constructive comments.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. ABB Group. Process Performance IE3 Aluminum Motors, 50/60 Hz. The Product Range for Global Customers,
Product Note. 2017. Available online: https://www.baldor.com/mvc/DownloadCenter/Files/9AKK107020
(accessed on 9 June 2020).
2. Almeida, A.; Ferreira, F.; Baoming, G. Beyond Induction Motors—Technology Trends to Move Up Efficiency.
IEEE Trans. Ind. Appl. 2014, 50, 2103–2114. [CrossRef]
3. Commission Regulation (EU) 2019/1781 of 1 October 2019 Laying Down Ecodesign Requirements for
Electric Motors and Variable Speed Drives Pursuant to Directive 2009/125/EC of the European Parliament
and of the Council, Amending Regulation (EC) No 641/2009 with Regard to Ecodesign Requirements
for Glandless Standalone Circulators and Glandless Circulators Integrated in Products and Repealing
Commission Regulation (EC) No 640/2009, Document 32019R1781. Available online: https://eur-lex.europa.
eu/legal-content/EN/TXT/PDF/?uri=CELEX:32019R1781&from=EN (accessed on 9 June 2020).
4. Doppelbauer, M. Update on IEC motor and converter standards. In Proceedings of the 6th International
Motor Summit for Energy Efficiency powered by Impact Energy, Motor Summit 2016, Zurich, Switzerland,
11–12 October 2016.
5. Annex to the Communication from the Commission to the European Parliament, the European Council,
the Council, the European Economic and Social Committee and the Committee of the Regions, The European
Green Deal, Brussels, 11.12.2019, COM(2019) 640 Final. Available online: https://ec.europa.eu/info/sites/info/
files/european-green-deal-communication-annex-roadmap_en.pdf (accessed on 9 June 2020).

15
Energies 2020, 13, 3546

6. WQuattro, Super Premium Efficiency Motor. WEG Group, Catalogue, Code 50025713, Revision 3 July 2017.
Available online: https://static.weg.net/medias/downloadcenter/h01/hfc/WEG-w22-quattro-european-market-
50025713-brochure-english-web.pdf (accessed on 9 June 2020).
7. Synchro Vert IE4 Super Premium Efficiency Motors, Bharat Bijlee, Catalogue. Available online: https:
//www.bharatbijlee.com/media/14228/synchrovert_catalogue.pdf (accessed on 5 May 2020).
8. Ismagilov, F.; Vavilov, V.; Gusakov, D. Line-Start Permanent Magnet Synchronous Motor for Aerospace
Application. In Proceedings of the 2018 IEEE International Conference on Electrical Systems for Aircraft,
Railway, Ship Propulsion and Road Vehicles and International Transportation Electrification Conference
(ESARS-ITEC 2018), Nottingham, UK, 7–9 November 2018; pp. 1–5. [CrossRef]
9. Sorgdrager, J.; Wang, R.-J.; Grobler, A. Multi-Objective Design of a Line-Start PM Motor Using the Taguchi
Method. IEEE Trans. Ind. Appl. 2018, 54, 4167–4176. [CrossRef]
10. Kurihara, K.; Rahman, M. High-Efficiency Line-Start Interior Permanent-Magnet Synchronous Motors.
IEEE Trans. Ind. Appl. 2004, 40, 789–796. [CrossRef]
11. Azari, M.; Mirsalim, M. Line-start permanent-magnet motor synchronisation capability improvement using
slotted solid rotor. IET Electr. Power Appl. 2013, 7, 462–469. [CrossRef]
12. Ghahfarokhi, M.; Aliabad, A.; Boroujeni, S.; Amiri, E.; Faradonbeh, V. Analytical modelling and optimization
of line start LSPM synchronous motors. IET Electr. Power Appl. 2020, 14, 398–408. [CrossRef]
13. Sethupathi, P.; Senthilnathan, N. Comparative analysis of line-start permanent magnet synchronous motor
and squirrel cage induction motor under customary power quality indices. Electr. Eng. 2020, 1–11. [CrossRef]
14. Dent, P. Rare earth elements and permanent magnets. J. Appl. Phys. 2012, 111, 07A721. [CrossRef]
15. Goss, J.; Popescu, M.; Staton, D. A Comparison of an Interior Permanent Magnet and Copper Rotor Induction
Motor in a Hybrid Electric Vehicle Application. In Proceedings of the IEEE International Electric Machines &
Drives Conference, EMDC, Chicago, IL, USA, 12–15 May 2013; pp. 12–15. [CrossRef]
16. Lima, I.; Filho, W. Rare Earth Industry; Elsevier: Amsterdam, The Netherlands, 2015.
17. Kazakbaev, V.; Prakht, V.; Dmitrievskii, V.; Ibrahim, M.N.; Oshurbekov, S.; Sarapulov, S. Efficiency Analysis of
Low Electric Power Drives Employing Induction and Synchronous Reluctance Motors in Pump Applications.
Energies 2019, 12, 1144. [CrossRef]
18. Moghaddam, R.; Magnussen, F.; Sadarangani, C. Theoretical and Experimental Reevaluation of Synchronous
Reluctance Machine. IEEE Trans. Ind. Electron. 2010, 57, 6–13. [CrossRef]
19. Kabir, M.; Husain, I. Application of a Multilayer AC Winding to Design Synchronous Reluctance Motors.
IEEE Trans. Ind. Appl. 2018, 54, 5941–5953. [CrossRef]
20. Pellegrino, G.; Cupertino, F.; Gerada, C. Automatic Design of Synchronous Reluctance Motors Focusing on
Barrier Shape Optimization. IEEE Trans. Ind. Appl. 2015, 51, 1465–1474. [CrossRef]
21. Kazakbaev, V.; Prakht, V.; Dmitrievskii, V.; Sarapulov, S.; Askerov, D. Comparison of power consumption
of synchronous reluctance and induction motor drives in a 0.75 kW pump unit. In Proceedings of the
2017 International Siberian Conference on Control and Communications (SIBCON), Kazakhstan, Astana,
29–30 June 2017; pp. 1–6. [CrossRef]
22. ABB Group. Low Voltage IE4 Synchronous Reluctance Motor and Drive Package for Pump and Fan Applications,
Catalogue. 2013. Available online: https://library.e.abb.com/public/21ee11b9fddfa677c1257bd500219300/
Catalog_IE4_SynRM_EN_06-2013_9AKK105828_lowres.pdf (accessed on 9 June 2020).
23. Jia, S.; Zhang, P.; Liang, D.; Dai, M.; Liu, J. Design of IE4 Level Synchronous Reluctance Machines with
Different Number of Poles. In Proceedings of the 2019 22nd International Conference on Electrical Machines
and Systems (ICEMS), Harbin, China, 11–14 August 2019; pp. 1–5. [CrossRef]
24. Dmitriveskii, V.; Prakht, V.; Kazakbaev, V. IE5 Energy-Efficiency Class Synchronous Reluctance Motor with
Fractional Slot Winding. IEEE Trans. Ind. Appl. 2019, 55, 4676–4684. [CrossRef]
25. Dmitrievskii, V.; Prakht, V.; Kazakbaev, V.; Oshurbekov, S.; Sokolov, I. Developing ultra premium efficiency
(IE5 class) magnet-free synchronous reluctance motor. In Proceedings of the 2016 6th International Electric
Drives Production Conference (EDPC), Nuremberg, Germany, 30 November–1 December 2016; pp. 2–7.
[CrossRef]
26. REEL SuPremE® –The IE5* Magnet-Free Synchronous Reluctance Motor, KSB Company, Catalogue. 2019.
Available online: https://www.ksb.com/blob/199782/10a90edea2eca4cb28afab3076d43df3/reel-supreme-
motor-brochure-en-dow-data.pdf (accessed on 8 July 2020).

16
Energies 2020, 13, 3546

27. SIMOTICS Reluctance Motors with SINAMICS Converters, Siemens AG, Catalog. 2015. Available online:
https://www.asberg.ru/upload/iblock/919/9195144e44c436fd94d723be0813ecfe.pdf (accessed on 9 June 2020).
28. Kersten, A.; Liu, Y.; Pehrman, D.; Thiringer, T. Rotor Design of Line-Start Synchronous Reluctance Machine
with Round Bars. IEEE Trans. Ind. Appl. 2019, 55. [CrossRef]
29. Liu, H.; Lee, J. Optimum Design of an IE4 Line-Start Synchronous Reluctance Motor Considering
Manufacturing Process Loss Effect. IEEE Trans. Ind. Electron. 2018, 65, 3104–3114. [CrossRef]
30. Kärkkäinen, H.; Aarniovuori, L.; Niemelä, M.; Pyrhönen, J.; Kolehmainen, J.; Känsäkangas, T.; Ikäheimo, J.
Direct-On-Line Synchronous Reluctance Motor Efficiency Verification with Calorimetric Measurements.
In Proceedings of the 2018 XIII International Conference on Electrical Machines (ICEM), Alexandroupoli,
Greece, 3–6 September 2018; pp. 171–177. [CrossRef]
31. Hu, Y.; Chen, B.; Xiao, Y.; Shi, J.; Li, L. Study on the Influence of Design and Optimization of Rotor Bars
on Parameters of a Line-Start Synchronous Reluctance Motor. IEEE Trans. Ind. Appl. 2020, 56, 1368–1376.
[CrossRef]
32. Gamba, M.; Armando, E.; Pellegrino, G.; Vagati, A.; Janjic, B.; Schaab, J. Line-start synchronous reluctance
motors: Design guidelines and testing via active inertia emulation. In Proceedings of the 2015 IEEE Energy
Conversion Congress and Exposition (ECCE), Montreal, QC, Canada, 20–24 September 2015; pp. 4820–4827.
[CrossRef]
33. Weatherization and Intergovernmental Programs Office, About the State Energy Program. Available online:
https://www.energy.gov/eere/wipo/about-state-energy-program (accessed on 9 June 2020).
34. Federal Council Aims for a Climate-Neutral Switzerland by 2050. Available online: https://www.admin.ch/
gov/en/start/documentation/media-releases.msg-id-76206.html (accessed on 9 June 2020).
35. Hopes of Limiting Global Warming? China and the Paris. Agreement on Climate Change. Available online:
https://journals.openedition.org/chinaperspectives/6924?file=1 (accessed on 9 June 2020).
36. Japan Net Zero Energy Building Roadmap. Available online: http://www.nzeb.in/wp-content/uploads/2015/
09/Japan-NZEB-Roadmap.pdf (accessed on 9 June 2020).
37. Ha, S.; Tae, S.; Kim, R. A Study on the Limitations of South Korea’s National Roadmap for Greenhouse Gas
Reduction by 2030 and Suggestions for Improvement. Sustainability 2019, 11, 3969. [CrossRef]
38. Shankar, V.; Umashankar, S.; Paramasivam, S.; Hanigovszki, N. A comprehensive review on energy efficiency
enhancement initiatives in centrifugal pumping system. Appl. Energy 2016, 181, 495–513. [CrossRef]
39. Stoffel, B. Assessing the Energy Efficiency of Pumps and Pump Units. In Background and Methodology; Elsevier:
Amsterdam, The Netherlands, 2015.
40. Goman, V.; Oshurbekov, S.; Kazakbaev, V.; Prakht, V.; Dmitrievskii, V. Energy Efficiency Analysis of
Fixed-Speed Pump Drives with Various Types of Motors. Appl. Sci. 2019, 9, 5295. [CrossRef]
41. Ahonen, T.; Orozco, S.; Ahola, J.; Tolvanen, J. Effect of electric motor efficiency and sizing on the energy
efficiency in pumping systems. In Proceedings of the 2016 18th European Conference on Power Electronics
and Applications (EPE’16 ECCE Europe), Karlsruhe, Germany, 5–9 September 2016; pp. 1–9. [CrossRef]
42. Rhyn, P.; Pretorius, J. Utilising high and premium efficiency three phase motors with VFDs in a public water
supply system. In Proceedings of the 2015 IEEE 5th International Conference on Power Engineering, Energy
and Electrical Drives (POWERENG), Riga, Latvia, 11–13 May 2015; pp. 497–502. [CrossRef]
43. Brinner, T.; McCoy, R.; Kopecky, T. Induction versus Permanent-Magnet Motors for Electric Submersible
Pump Field and Laboratory Comparisons. IEEE Trans. Ind. Appl. 2014, 50, 174–181. [CrossRef]
44. Ogunjuyigbe, A.; Jimoh, A.; Nicolae, D.; Obe, E. Analysis of synchronous reluctance machine with
magnetically coupled three-phase windings and reactive power compensation. IET Electr. Power Appl. 2010,
4, 291–303. [CrossRef]
45. Obe, E. Steady-state performance of a line-start synchronous reluctance motor with capacitive assistance.
Electr. Power Syst. Res. 2010, 80, 1240–1246. [CrossRef]
46. Gamba, M.; Pellegrino, G.; Vagati, A.; Villata, F. Design of a line-start synchronous reluctance motor.
In Proceedings of the 2013 International Electric Machines & Drives Conference, Chicago, IL, USA, 12–15 May 2013;
pp. 648–655. [CrossRef]
47. Wquattro 4 kW 4P 112M 3Ph 400/690 V 50 Hz IC411-TEFC-B3T. Product Details. Available online: https://
www.weg.net/catalog/weg/RS/en/Electric-Motors/Special-Application-Motors/Permanent-Magnet-Motors/
Line-Start-PM-Motors/Wquattro-4-kW-4P-112M-3Ph-400-690-V-50-Hz-IC411---TEFC---B3T/p/13009391
(accessed on 9 June 2020).

17
Energies 2020, 13, 3546

48. W22 IE3 4 kW 4P 112M 3Ph 220-240/380-415//460 V 50 Hz IC411-TEFC-B14T. Product Details. Available online:
https://www.weg.net/catalog/weg/RS/en/Electric-Motors/Low-Voltage-IEC-Motors/General-Purpose-
ODP-TEFC/Cast-Iron-TEFC-General-Purpose/W22---Cast-Iron-TEFC-General-Purpose/W22-IE3/W22-
IE3-4-kW-4P-112M-3Ph-220-240-380-415-460-V-50-Hz-IC411---TEFC---B14T/p/14640586 (accessed on
9 June 2020).
49. Matine, A.; Bonnard, C.-H.; Blavette, A.; Bourguet, S.; Rongère, F.; Kovaltchouk, T.; Schaeffer, E. Optimal
sizing of submarine cables from an electro-thermal perspective. In Proceedings of the European Wave and
Tidal, Conference (EWTEC), Cork, Ireland, 27 August–1 September 2017; Available online: https://hal.archives-
ouvertes.fr/hal-01754060 (accessed on 9 June 2020).
50. NM; NMS. Close Coupled Centrifugal Pumps with Flanged Connections; Catalogue; Calpeda. 2018.
Available online: https://www.calpeda.com/system/pdf/catalogue_en_50hz.pdf (accessed on 5 May 2020).
51. IE4 DOLSynRM Concept Introduction, SynRM Benefits Available for Fixed Speed Applications.
Tailored High Efficiency Motors for OEMs, Document Number: 9AKK106542 EN 04-2015. Available
online: https://search.abb.com/library/Download.aspx?DocumentID=9AKK106542&LanguageCode=en&
DocumentPartId=&Action=Launch (accessed on 9 June 2020).
52. Extended Product Approach for Pumps, Europump. 2014. Available online: http://europump.net/uploads/
Extended%20Product%20Approach%20for%20Pumps%20-%20A%20Europump%20guide%20(27OCT2014)
.pdf (accessed on 9 June 2020).
53. Zhou, J.-H.; Chen, C.; Zhang, X.-W.; Chen, Y.-A. Reducing voltage energy-saving control method of induction
motor. In Proceedings of the 2013 International Conference on Electrical Machines and Systems (ICEMS),
Busan, Korea, 26–29 October 2013; pp. 2159–2162. [CrossRef]
54. Kirar, M.; Aginhotri, G. Cable sizing and effects of cable length on dynamic performance of induction motor.
In Proceedings of the 2012 IEEE Fifth Power India Conference, Murthal, India, 19–22 December 2012; pp. 1–6.
[CrossRef]
55. Electrical Installations in Buildings—Part 5–52: Selection and Erection of Electrical Equipment—Wiring Systems,
Is the IEC Standard Governing Cable Sizing; IEC 60364-5-52; IEC: Geneva, Switzerland, 2009; Available online:
https://webstore.iec.ch/publication/1878 (accessed on 8 July 2020).
56. Eurostat Data for the Industrial Consumers in Germany. Available online: http://ec.europa.eu/eurostat/
statistics-explained/index.php/Electricity_price_statistics#Electricity_prices_for_industrial_consumers
(accessed on 9 June 2020).
57. Pump Life Cycle Costs: A Guide to LCC Analysis for Pumping Systems, Executive Summary; Hydraulic Institute:
Parsippany, NJ, USA; Europump: Brussels, Belgium; Office of Industrial Technologies Energy Efficiency and
Renewable Energy, U.S. Department of Energy: Washington, DC, USA, 2001; pp. 1–19. Available online:
https://www.energy.gov/sites/prod/files/2014/05/f16/pumplcc_1001.pdf (accessed on 8 July 2020).
58. Waghmode, L.; Sahasrabudhe, A. A comparative study of life cycle cost analysis of pumps. In Proceedings
of the ASME 2010 International Design Engineering Technical Conferences and Computers and Information
in Engineering Conference (ASME 2010), Montreal, QC, Canada, 15–18 August 2010; pp. 491–500. [CrossRef]

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

18
energies
Article
Simulation-Based Coyote Optimization Algorithm to
Determine Gains of PI Controller for Enhancing
the Performance of Solar PV Water-Pumping System
Jouda Arfaoui 1 , Hegazy Rezk 2,3, *, Mujahed Al-Dhaifallah 4, *, Mohamed N. Ibrahim 5,6,7
and Mami Abdelkader 8
1 National School of Engineering of Tunis, University of Tunis ELMANAR, BP 37, Tunis 1002, Tunisia;
[email protected]
2 College of Engineering at Wadi Addawaser, Prince Sattam Bin Abdulaziz University,
Al-Kharj 11991, Saudi Arabia
3 Electrical Engineering Deprtment, Faculty of Engineering, Minia University, Al Minya 61519, Egypt
4 Systems Engineering Department, King Fahd University of Petroleum & Minerals,
Dhahran 31261, Saudi Arabia
5 Department of Electromechanical, Systems and Metal Engineering, Ghent University, 9000 Ghent, Belgium;
[email protected]
6 FlandersMake@UGent—Corelab EEDT-MP, 3001 Leuven, Belgium
7 Electrical Engineering Department, Kafrelshiekh University, Kafr el-Sheikh 33511, Egypt
8 Department of Physics, Faculty of Sciences, University Tunis El Manar, BP 37, Tunis 1002, Tunisia;
[email protected]
* Correspondence: [email protected] (H.R.); [email protected] (M.A.-D.)

Received: 4 June 2020; Accepted: 27 August 2020; Published: 31 August 2020

Abstract: In this study, a simulation-based coyote optimization algorithm (COA) to identify the gains
of PI to ameliorate the water-pumping system performance fed from the photovoltaic system is
presented. The aim is to develop a stand-alone water-pumping system powered by solar energy,
i.e., without the need of electric power from the utility grid. The voltage of the DC bus was adopted as
a good candidate to guarantee the extraction of the maximum power under partial shading conditions.
In such a system, two proportional-integral (PI) controllers, at least, are necessary. The adjustment of
(Proportional-Integral) controllers are always carried out by classical and tiresome trials and errors
techniques which becomes a hard task and time-consuming. In order to overcome this problem,
an optimization problem was reformulated and modeled under functional time-domain constraints,
aiming at tuning these decision variables. For achieving the desired operational characteristics of
the PV water-pumping system for both rotor speed and DC-link voltage, simultaneously, the proposed
COA algorithm is adopted. It is carried out through resolving a multiobjective optimization problem
employing the weighted-sum technique. Inspired on the Canis latrans species, the COA algorithm
is successfully investigated to resolve such a problem by taking into account some constraints in
terms of time-domain performance as well as producing the maximum power from the photovoltaic
generation system. To assess the efficiency of the suggested COA method, the classical Ziegler–Nichols
and trial–error tuning methods for the DC-link voltage and rotor speed dynamics, were compared.
The main outcomes ensured the effectiveness and superiority of the COA algorithm. Compared to
the other reported techniques, it is superior in terms of convergence rapidity and solution qualities.

Keywords: simulation-based optimization; coyote optimization algorithm; water pumping;


energy efficiency

Energies 2020, 13, 4473; doi:10.3390/en13174473 www.mdpi.com/journal/energies

19
Energies 2020, 13, 4473

1. Introduction
Thanks to the huge availability of the sun supply in comparison with other energy sources,
the solar PV systems seem one of the most encouraging applications of renewable energy today
and in the forthcoming years. Moreover, the cost of the PV modules is reducing increasing a result of
the progress in the manufacturing technology of the solar modules [1]. Standalone and grid-connected
systems are the two main types of solar systems. Indeed, Sun availability is taking as one of the most
important motivation for depending on the solar system as a promising solution. As being a type of
renewable energy resources, the solar PV has gained a growing importance for uses in the electric
power domain which has several advantages like (1) being harmless energy, (2) being suitable for
isolated sites and (3) being cheap and easy maintenance.
Indeed, the efficiency and cost of the PV pumping system are not attractive yet to spread widely,
but, many literature works are devoted to improving the total effectiveness and reducing the total cost
of the solar pumping generation system [1,2].
A major drawback of using this sort of vector control lies in the fact that the effectiveness of
the water-pumping system is in the heavy dependence on the appropriate adjustment of the PI
parameters. Commonly, the set of these parameters is based on the classical and tiresome trials
and error methods [3–6].
Adjustment method turned out to be complex and time-consuming. Moreover, the conventional
PI tuning approach such as the symmetrical optimum [7], Ziegler–Nichols [8], Tyreus–Luyben [9]
and Cohen–Coon [10] techniques requires that the designer has a good insight about the characteristics
and dynamics of the controlled system. Hence, for the sake of overcoming these drawbacks, a systematic
process to identify best gains of PI regulators has been provided in [3] and the modern optimization is
considered as an auspicious solution.
In [11] the authors considered two optimization metaheuristics: Ant colony optimization (ACO)
and differential evolution (DE). These strategies are reported to provide an optimized adjustment (PI)
regulators in the direct torque control–space vector modulation (DTC-SVM) control loops, like rotor
speed, electromagnetic torque, stator flux linkage and estimation of the linkage stator flux. Compared
to other metaheuristics methods, those considered approaches require few tuning parameters to
provide a fast convergence rate. Simulation and practical experimental on three phase induction
motor (TIM) controlled by DTC-SVM are carried out. On the other hand, classical procedures such as
frequency and root locus have been applied.
In [12], the study reported a novel optimization algorithm that combines all these three techniques:
genetic algorithm (GA), artificial neural networks (ANN) and fuzzy logic, which was named as GNFPID.
This technique aims to outperform the performance of the used PID controller. In [13], a genetic
algorithm is applied as the adequate adjustment method of PI controllers applied to DTC drive for
doubly fed asynchronous machine. Following in the same strategy, particle-swarm optimization (PSO),
bat algorithm (BA) and genetic algorithm (GA) are considered for the adequate tuning of the gains of
PI for brushless DC motor in [14].
Metaheuristics are an iterative processes for solving different optimization problems by taking
into account simple trial operations [15,16]. Particle-swarm optimization [17], genetic algorithm [18],
Imperialist competitive algorithm [19], ant colony optimization [20,21], differential evolution [22,23]
are commonly treated. In [24], water cycle algorithm was adopted to adjust the parameters of PI
controller for speed control and reactive power control loops in RSC (rotor side converter) and DC-link
regulation voltage loop in GSC (grid side converter).
In the same way, [25] has referred to a thermal exchange optimization (TEO), this metaheuristic
method was applied for the tuning of PI controller gains for the external loops in the conventional
vector control strategy of a doubly fed induction generator based on wind turbine system. For assessing
the efficiency and superiority of the proposed control strategy, other well-known metaheuristic
procedures such as grasshopper optimization algorithm (GOA), particle-swarm optimization (PSO),

20
Energies 2020, 13, 4473

water cycle algorithm (WCA) and harmony search algorithm (HSA), are taken into account for
comparison purposes.
To best of our knowledge, for first time, the coyote optimization (COA) algorithm is used to
identify the parameters of PI controller to improve the performance of solar PV water-pumping system
under partial shading condition. Therefore, in this study, COA algorithm is investigated and further,
to prove the superiority of the suggested COA algorithm, the obtained results are compared with
the classical Ziegler–Nichols method, trial–error tuning method and PSO algorithm.
The remainder of this study is arranged as follows: the mathematical model of the solar
photovoltaic water-pumping system (SPWPS) is described in Section 2. Section 3 is devoted to
the formulation of the outer-loops PI controllers’ design problem given as a constrained optimization
problem. The proposed COA method is given in detail, aimed to resolve such a problem. In Section 4,
the demonstrative results of the COA-tuned PI controllers tuning are carried out. Concluding remarks
are summarized in Section 5.

2. Modeling of the Solar Photovoltaic Water-Pumping System


Referring to the solar pumping system, several control methods were developed to provide an
effective functioning. These include speed control, distinct maximum power point tracking (MPPT)
techniques, voltage control and frequency regulation under load variation [26].
This part studies and simulates a new structure of a solar pumping system. The overall structure
of the proposed PV pumping system is described, as well as its various subsystems.
Vector control also called field-oriented control, is a method that permits independent control
of the flux and torque components. It is equipped with high-performance drives. In this section,
the interest is focused on the type of control.
The proposed system includes the following parts: PV panel, three-phase voltage inverter, DC–DC
boost converter and induction engine-driven pump unit.
A schematic diagram of the new structure including indirect field oriented control (IFOC)
and MPPT strategies, is illustrated in Figure 1. Concerning the control of this sort of system, different
techniques were considered for effective functioning in terms of speed control, voltage control, distinct
MPPT techniques and frequency regulation under load variations. The most important parts of
this system are indicated in Figure 1 and are recapitulated as follows:

(1) Shaded PV modules represent the power supply source of the induction engine via a three-phase
voltage source inverter (VSI);
(2) A hysteresis pulse-width modulation (PWM) based current control;
(3) A boost converter for forcing the PV panel to operate at the MPP under different partial shading
conditions [27];
(4) Flux weakening component is required to produce the reference current (Ids ∗ ) and the speed

controller output gives (Iqs );
(5) Vector control of asynchronous machine for driving centrifugal pump [28].

21
Energies 2020, 13, 4473

Figure 1. Scheme diagram of the proposed system.

The reference speed is a measure of the generated PV power. The PI controllers’ design
and the suitable values of the gains regulators (Kps , Kis , Kpdc and Kidc ) are typically tuned employing
trials and errors technique, as mentioned before. Regarding the control design stage of VSI, this hard
and challenging task turns into more difficult and time-consuming. A recent global meta-heuristic
algorithm is an effective solution to tackle this matter.

2.1. Model of the Induction Machine


In this context, the Park transformation is adopted when modeling the asynchronous machine to
omit the inductance variation with time [29].

dIds Lm dϕdr
Vds = Rs Ids + σLs + − ωs σLs Iqs (1)
dt Lr dt

dIqs Lm
Vqs = Rs Iqs + σLs + ωs ϕ + ωs σLs Ids (2)
dt Lr dr
Rr dϕdr Lm
Vdr = 0 = ϕ + − Rr Ids (3)
Lr dr dt Lr
Lm
Vqr = 0 = ωsl ϕdr − Rr Iqs (4)
Lr
where
Lm 2
σ = 1−
Ls Lr
Vds , Vqs , Vdr and Vqr indicate the stator and the rotor voltages, respectively.

22
Energies 2020, 13, 4473

Ids and Iqs represent the stator currents.

ϕdr = Lm Imdr (5)

Lr dImdr
Ids = Imdr + (6)
Rr dt
Ids = (1 + p·τr )Imdr (7)

where
Lr
τr =
Rr
In steady states conditions
Ids = Imdr (8)

The instantaneous value of the angle θs is determined by this equation:



θs = ωs dt (9)

 
θs = ωs dt = (ωsl + ωr )dt (10)

The sliding speed ωsl can be calculated using this equation [29]:

Lm Rr Lm Iqs Iqs
ωsl = Iqs = = (11)
Lr ϕdr τr ϕdr τr Imdr

In steady states conditions (Imdr = Ids ) and thus

Iqs
ωsl = (12)
τr Ids

Equations (13) and (14) are needed to determine the reference flux and reference stator current,
respectively [29]:

ϕrn si |ωm | ≤ ωmn
ϕ∗dr = ϕrn ωmn (13)
|ωm | si |ωm | > ωmn
where ϕrn denotes nominal value of flux.
ωmn denotes nominal value of mechanical speed


ϕ∗dr
Ids = (14)
Lm

Lm : is the mutual inductance


∗ 3 PLm ∗
Iqs = Te∗ / ϕ (15)
2 2Lr dr
The correlation process between the torque and the speed of the pump model is described as
follows [30]:
Tr = a1 + a2 ·ωm 2 (16)

where
a1 : constant;
a2 : constant;
Tr : Torque of pump;
wm : motor speed.

23
Energies 2020, 13, 4473

The power of the pump has a cubic relationship with the motor speed wm , thus driving
the centrifugal pumps [31,32].
To evaluate one of the reference speed components, an affinity law shall be applied for this purpose.
Under different solar irradiation, the induction machine operates and seeks to determine the centrifugal
pump flow rates. Moreover, to optimize the produced PV power, a MPPT control was investigated.
It enables the extraction of the global MPP. The generated power serves to determine the first component
of the reference speed as follows [33]:
 
Ppv 1/3
ωm1 = (17)
k
The PV array power is converted in terms of speed via the constant k.
The second component of the reference speed is evaluated by means of the DC-link voltage
controller. In this system, a comparison between the measured bus voltage (Vdc ) and the desired bus
∗ ), is carried out, which conducts to a voltage error (ΔV ), described as [34]:
voltage (Vdc dc


ΔVdc (n) = Vdc (n) − Vdc (n) (18)

The output speed can be obtained by the following equation [34]:




ωm2 (n) = ωm2 (n − 1) + Kpdc ΔVdc (n)− ΔVdc (n − 1) + Kidc ·ΔVdc (n) (19)

where Kpdc and Kidc are, respectively the proportional and integral parameters of the DC-link-
voltage controller.
The reference speed of the induction motor is obtained according to the following equation [33]:

ω∗m = ωm1 + ωm2 (20)

where ω∗m is the reference speed and wm1 is the first part of reference speed, determined by means of
Equation (17).
wm2 is the second part of reference speed, determined by means of Equation (19).

2.2. Pump Model


The mechanical model of the induction motor actuating the centrifugal pump is provided by
the following formula [35]:
dωm
Te − Tr − f ωm = J (21)
dt
Tr and Te represent the load torque and the electromagnetic torque, respectively, J is the moment
of inertia and f is the damping coefficient which links the friction torque Cf to the motor speed by
relation (C f = − f ωm ).
The following equation used to represent the pump torque [33]:

Tr = Ap ω2m (22)

where Ap denotes torque constant and estimates as following [33]:

Pn
Ap = (23)
ω3mn

where Pn and wmn indicate rated power and nominal value of mechanical speed, respectively.

24
Energies 2020, 13, 4473

2.3. Three-Phase Voltage Inverter


To achieve an optimum control of the power supply between the PV source and the asynchronous
engine, a voltage source inverter is required as a crucial part. The employed inverter consists of three
independent arms, as illustrated in Figure 2. DC–AC converters are dealing with in this section, which
already applied in the solar energy sector as well as in the three-phase voltage inverters. The state
of the switches (kn ) has a considerable impact on the inverter modeling. The produced voltages are
provided by the following matrix form [36]:
⎡ ⎤ ⎡ ⎤⎡ ⎤
⎢⎢ V1 ⎥⎥ ⎢ 2 −1 −1 ⎥⎥⎢⎢ S1 ⎥⎥
⎢⎢ ⎥ V ⎢⎢ ⎥⎥⎢⎢ ⎥
⎢⎢ V2 ⎥⎥⎥ = pv ⎢⎢⎢ −1 2 −1 ⎥⎥⎢⎢ S2 ⎥⎥⎥ (24)
⎢⎢ ⎥⎥ 3 ⎢⎢⎣ ⎥⎥⎢⎢ ⎥⎥
⎣ ⎦ ⎦⎣ ⎦
V3 −1 −1 2 S3

/Ɖǀ

<ϰ <ϱ <ϲ



sƉǀ Ϯ
Ϯ /ƐĂ

sϮ /Ɛď
K
sϯ /ƐĐ
<ϭ <Ϯ <ϯ
sƉǀ 
Ϯ Ϯ

^ϭ ^Ϯ ^ϯ

Figure 2. Three-phase voltage inverter.

The current provided by the DC source obeys to;

Ipv = S1 Isa + S2 Isb + S3 Isc + Ic (25)

A PWM hysteresis strategy is employed to control these power converters.

2.4. DC–DC Boost Converter


An ideal scheme of the DC–DC boost converter is illustrated in Figure 3. The relation of
proportionality between the input voltage (E) and the output voltage, is described as follows:

E
Vs = (26)
1−u
where u denotes the duty cycle.

25
Energies 2020, 13, 4473

>
ŝ>

ŝŽĚĞ
ϭ
ʅ

н
Z͕ sƐ
 DŽƐĨĞƚ

Figure 3. Scheme of DC converter.

The DC converter can be integrated with MPPT strategy to extract the maximum power from
the PV array.

2.5. Hysteresis Based PWM Current Controller


In most studies, the PWM technique is the most common strategy since it perfectly adapts with
the use of the voltage inverter. It is recommended also in case the vector control is adopted.
The principle of this command as presented in Figure 4 is to generate a PWM signal based directly
on the variable to be controlled, through decisions based on the on–off control. In addition, it consists
of maintaining the variation of the measuring currents within a band at a self-adjusting width, centered
on the reference currents. The hysteresis comparator of any PWM structure receives as an input signal:
the difference between the reference and measurement currents in order to produce at the output
the control signal for the power switches [36]. In fact, this control is simple to operate as it does not
require precise knowledge of the machine but suffers from the lack of control of the switching frequency.

ŽŶƚƌŽůŽĨŬ

ZĞĨĞƌĞŶĐĞĐƵƌƌĞŶƚ н

Ͳ ŽŶƚƌŽůŽĨŬ͛

+\VWHUHVLVFRPSDUDWRU

/ŶǀĞƌƚŝŶŐŐĂƚĞ
DĞĂƐƵƌĞĚĐƵƌƌĞŶƚ

Figure 4. Schematic diagram of the hysteresis pulse-width modulation (PWM) control.

3. COA Design for Optimal PI Parameters Tuning


The coyote optimization algorithm (COA) was recently proposed by Pierezan and Coelho [37].
The core idea of COA optimizer is based on Canis latrans species that reside mainly North America [37].
The algorithm is adapted to consider the social organization of the agents called coyotes and it has
been serving with a different algorithmic construction. The greatest advantage of this method is to
preserve a balance between the exploitation and exploration mechanisms during that optimizing
process. By contrast with the gray wolf optimization (GWO), the COA does not center on how these
dominant norms have been followed by these animals and the social hierarchy. Additionally, the COA
depends not only on the hunting preys which take place in the GWO, but also on the social structure
and regular experience interchange which are carried out by the coyotes. They are distinguished by
cooperative functionalities as they head toward the prey in the close chain while they have a strong

26
Energies 2020, 13, 4473

sense of smell which makes it possible to identify the location of the prey. Regarding the hunting
process, coyotes attack in groups, this action forces the agents to update their positions to improve
them. When coyotes’ striking their opponents, they are well-prepared with a probability of threat
and move away from their current position by an excessive random distance. Further, the COA method
is initialized by an initializing population of coyotes whose size can be defined as the multiplication
of packs and coyotes. The number of agents called coyotes in each pack is supposed to be equal
and constant. In that regard, to facilitate the users’ understanding, each coyote can be considered as a
candidate solution for the optimization problem and its social condition presents the fitness function.
The COA optimization algorithm started by a population of coyotes that are created randomly within
the search space. More details about physical meaning and mathematical representation can be found
in Ref [37,38]. Figure 5 presents the optimization process of COA.

,QLWLDOL]DWLRQ 6WDUW

'HILQHLQSXWSDUDPHWHUV
DQGLQLWLDOL]DWLRQ
&R\RWHORRS
(VWLPDWHWKHFRUUHVSRQGLQJILWQHVVIXQFWLRQ 
t 
3HUIRUPELUWKDQGGHDWK
p 
SURFHVVLQVLGHWKHSDFN

,GHQWLI\WKHFR\RWHUHSUHVHQWV
WKHDOSKDRIWKHSDFN QR Last pack?
\HV
(VWLPDWHWKHFXOWXUDO
WHQGHQF\RIWKHSDFN 3HUIRUPSDFN¶V
WUDQVLWLRQV
c 
0RGLI\DJHVRI
(VWLPDWHWKHQHZVRFLDO FR\RWHV
FRQGLWLRQ

0RGLI\WKHVRFLDO 1H[WLWHUDWLRQ
FRQGLWLRQ LWHU LWHU LWHU>LWHUPD[
QR
\HV
2XWSXWWKHJOREDO
c c
QR Last Coyote?  EHVWFR\RWH
\HV
3DFNORRS (QG

Figure 5. Optimization process of coyote optimization algorithm (COA).

Regarding the PI design control, the suitable parameters of this type of controllers are usually
tuned by trials and error techniques [3–6]. This non-systematic process turns into more complicated
and into more time-consuming, particularly for the complex systems’ design. Accordingly, the idea
of transforming the identification process of these parameters (Kps , Kis , Kpdc and Kidc ) to an
optimization problem, is a promising alternative. By means of some advanced meta-heuristic
algorithms, such optimization problems can be effectively managed [3]. Within the framework

27
Energies 2020, 13, 4473

provided, two interactive PI regulators for speed and DC-link voltage loops are being introduced.
These PI regulators are demanded to achieve an optimized value of this control using such COA
algorithm. Figure 6 provides the suggested tuning process of PI controllers for the solar photovoltaic
water-pumping system (SPVWPS) based on optimization approach.

Figure 6. Proposed metaheuristics-tuned proportional-integral (PI) controllers for the solar photovoltaic
water-pumping system (SPVWPS).

During the optimization process, the gains of PI controllers for the DC-link voltage and speed
loops are considered as the decision variables of the problem, these parameters are denoted as follows:
 T
x = Kpdc , Kidc , Kps , Kis ∈ S ⊆ R4+ (27)

where Kpdc and Kidc indicate the proportional and integral gains of DC-link regulator and Kps and Kis
denote the proportional and integral gains of speed regulator.
Under time-domain, such constraints like rise time, maximum overshoot, settling time
and steady-state error of the system step-response, the performance criterion is minimized according
to the suitable parameters of the problem [3].
The tuning process based on optimization is defined as follows:


⎪ minimize gi (x)

⎪  T



⎪ x = Kpdc , Kidc , Kps , Kis ∈ S ⊆ R4+



⎪ subject to : (28)



⎪ h1 (x) = δVdc − δmax ≤0

⎪ Vdc

⎩ h2 (x) = δspeed − δmax
speed
≤0

In formulating the optimization problem, cost function are outlined as follows:

∞  

gIAE,i (x) = ei (x, t)dt (29)
0

Which are adopted the Integral of Absolute error (IAE) and maximum overshoot (MO) criteria,
S = x ∈ R4+ , xlow ≤ x ≤ xup is the initial search space.

28
Energies 2020, 13, 4473

Under such circumstances and to show superior performance, some constraints had to be
considered in the optimization framework, h j : R4+ → R, j ∈ {1, 2} presented the inequality constraints
which are considered for this optimization process.
The terms δdc and δspeed are the overshoot of the DC-link voltage and rotor speed, respectively.
The optimization process described in Equation (28) is considered as a multi-objective optimization
type; an aggregate of all costs functions to form one single objective function is determined by means
of an aggregation function [3].
2
g(x) = ωi gi (x) (30)
i=1

where wi > 0, it denotes the weighting coefficients of the aggregation cost function. The objective
function befitting for COA-tuned PI regulators for SPVWPS under shading condition, reformulated as
follows:  +∞    
g(x) = (ω1 w∗r − wr  + ω2 v∗dc − vdc )dt (31)
0

4. Results and Discussion


The considered system in this research work contains a three-phase induction motor drive of
2200 W, 230 V, used to feed the water-pumping system. The main source is PV array with rating of
2400 W. The PV array composed of 234 PV module. The number of series PV modules interconnected is
series is 18 modules whereas number of parallel string is 13 strings. The open circuit voltage and short
circuit current of PV modules are 21.6 V and 0.64 A, respectively whereas the voltage and current at
maximum power point are 17.6 V and 0.58 A. The case study considered the partial shading condition,
three different level of the solar radiation; 1.0 kW/m2 , 0.8 kW/m2 and 0.5 kW/m2 are subjected to
the PV array.
In order to tackle the problem of time-consuming, the PI control tune by COA is performed, which
further highlights the contribution of the proposed adjustment approach based on metaheuristic versus
the given classical Ziegler–Nichols and Trial and Error method.
Table 1 provides a summary of the obtained gains of the PI controllers for the classical based
methods and the proposed optimization meta-heuristic called the coyote optimization algorithm (COA).
Furthermore, a comparison is carried out between the obtained results for the proposed COA-tuned PI
controllers and those obtained by the well-known Ziegler–Nichols and Trial and Error method for
the PI controllers’ design of DC-link voltage (Table 2 and Figure 7).

Table 1. Optimized PI controller gains.

Algorithms
Parameter Trial and Error Ziegler–Nichols COA
Kps 0.2499 4.78577 1.8706
Kis 16.9730 229.6941 3.8149
Kpdc 1 0.3052 2.7748
Kidc 500 62.2047 1.3374

Table 2. Output performance under different methods.

Methods δ (%) tr (sec) ts (sec) Ess


Trial–Error 20 0.1238 0.26 0.006
Ziegler–Nichols 5.1 0.0279 0.2 0.004
PSO-Method 0 0.0332 0.987 0.03
COA-Method 0 0.015 0.151 0.001

29
Energies 2020, 13, 4473

Figure 7. DC-link voltage under different methods.

According to these results, COA-tuned PI controllers with the IAE criteria, outperforms these
considered classical based methods in terms of settling time, overshoot and steady-state error indices.
Indeed, the DC-link voltage performance using COA for designing the PI controller has a negligible
steady-state error and also a negligible overshoot and it provides a shorter rise time than that in
the case of a conventional adjustment method which presents a higher overshoot at the transient
operation. For such reformulated PI design issues and in terms of offering the quality of solutions
and non-premature convergence as well as fastness, the COA algorithm outperforms all other methods.
To reinforce the effectiveness of the proposed COA method, the homologous PSO algorithm is
considered for comparison purposes.
Figure 8 represents the dynamic response of the DC-link voltage based on the PI parameters design
by COA and PSO. It can be observed from this figure, that COA can maintain the DC-link voltage
to constant value faster than that of the other meta-heuristic (PSO). This confirms the higher global
convergence of COA due to the impact of exploration and exploitation mechanisms that are developed
by that algorithm. From Figure 8, the tracking dynamic indicates a considerable oscillating using
the PSO algorithm for adjustment PI controller design and the COA tuned PI controller enables us to
reach rapidly the steady-state. To confirm the reliability of COA compared to PSO, both algorithms
executed 30 times. The population size and maximum of iterations are 20 and 50, respectively.
The results of statistical comparison are presented in Table 3.

30
Energies 2020, 13, 4473

Figure 8. DC-link voltage responses using particle-swarm optimization (PSO) and COA.

Table 3. Statistical comparison between PSO and COA.

Algorithms Best Mean Worst Median STD


PSO 3.4528 6.5660 9.8254 6.2777 2.23
COA 2.1696 5.3835 8.5281 5.8741 2.14

The detailed performance (power, speed, DC-link voltage, load torque, electromagnetic torque,
etc. . . . .) of the PV generation system, obtained with COA algorithm used for tuning PI controllers
design of the DC-link voltage and rotor speed, are illustrated in Figure 9.

Figure 9. Performance of SPVWPS under partial shading condition using COA algorithm tuned PI
controller of the DC link voltage and rotor speed. (a) power; (b) DC voltage; (c) speed; (d) electromagnetic
torque; (e) load torque and (f) stator currents.

31
Energies 2020, 13, 4473

Considering Figure 9, it can be observed that the desired and the detected values of the rotor speed
converge simultaneously to the nominal value that demonstrate the ameliorated efficiency of the speed
control. Considering the heavy dependence between the rotor speed components and the load torque,
the water spilling out of the pump was reduced due to sudden torque variation in the suggested
regulation method. Figure 9c illustrates the rotor speed behavior of the PV system due to the increased
proportionally of the load torque, however, it achieves steady-state shortly at 0.2 s.
However, the proposed system uses an MPPT algorithm that provides the reference speed as an
output signal, to be fed into indirect-field oriented control (IFOC) strategy. It is worth noting that
the reference and measured values of the DC bus voltage converge towards the desired value, which
has demonstrated an enhanced efficiency of the voltage regulation. The reference value (set-point) of
the DC bus voltage must be constant and its value was selected to 388 V. The outer PI controller loop is
adopted to keep the DC bus voltage constant. The actual value of the DC bus voltage is the value that
is measured by the voltage sensor.
In other words, this Set point (reference speed), which must be applied to the machine, comes
from a new maximum power tracking algorithm based on the PSO technique (PSO-MPPT) and a pump
affinity law (Equation (16)). This process results in an optimization of photovoltaic power under partial
shading conditions. Further, it can be seen that the load torque and the rotor speed follow the PV
power. This is found in Figure 9. It was found that the inaccuracy of the speed has an important effect
on the efficiency and reliability of the pump drive (The relationship of proportionality was indicated in
Equation (16)).
In sum, the COA algorithm can identify the gains of PI controllers of DC link voltage and rotor
speed in a reduced computational time and under functional constraints compared to the reported
classical techniques which usually provide local solutions for the control problem with constraints.

5. Conclusions
The optimal gains of PI controllers were determined using coyote optimization algorithm (COA)
in order to enhance the effectiveness of the solar photovoltaic water-pumping system. In fact,
the controllers’ adjustment is reformulated as a constraint optimization problem under the functional
time domain. A comparative study of the dynamic response of the DC-link voltage and rotor speed for
the COA-tuning PI controllers is carried out with the well-known classical Ziegler–Nichols and Trial
and Error tuning procedures. Similarly, the proposed meta-heuristic COA succeeds in the tune of PI
controllers for DC-link voltage and rotor speed more effectively than the considered meta-heuristic
PSO algorithm. The COA provides good responses of the DC-link voltage and rotor speed, during
the transient and steady state cycle. The main outcomes confirmed the performance superiority of
the developed COA-tuned PI controllers, in terms of the rapidity of convergence and solutions quality.
To reduce the cost of the controller, a senseless-MPPT will be considered in the future work.

Author Contributions: Conceptualization, J.A., M.N.I. and H.R.; methodology, J.A., M.N.I and H.R.; software,
J.A., H.R. and M.A.-D.; formal analysis, J.A., M.N.I., M.A. and H.R.; investigation, J.A., M.A.-D., M.A.; resources,
J.A, M.A.-D. and H.R.; writing—original draft preparation, J.A., M.N.I., M.A.-D., M.A. and H.R.; writing—review
and editing, J.A., M.N.I., M.A.-D., M.A. and H.R. All authors contributed equally to this work. All authors have
read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Conflicts of Interest: The authors declare no conflict of interest.Abbreviations

Abbreviations
ACO Ant colony optimization
ANN Artificial neural network
BA Bat algorithm
COA Coyote optimization algorithm
DC Direct current

32
Energies 2020, 13, 4473

AC Alternating current
DE Differential evolution
DTC-SVM Direct torque control–space vector modulation
GA Genetic algorithm
GNFPID Genetic neural fuzzy proportional integral derivative
GOA Grasshopper optimization algorithm
GSC Grid side converter
GWO Gray wolf optimization
HSA Harmony search algorithm
IAE Integral absolute error
IFOC Indirect field oriented control
MO Maximum overshoot
MPP Maximum power point
MPPT Maximum power point tracking
PI Proportional integral
PID Proportional integral derivative
PSO Particle-swarm optimization
PV Photovoltaic
PWM Pulse-width modulation
TIM Three phase induction motor
RSC Rotor side converter
SPVWPS Solar photovoltaic water-pumping system
TEO Thermal exchange optimization
VSI Voltage source inverter
WCA Water cycle algorithm
Symbols
f Damping coefficient
Ids ,Iqs Stator currents
Lm magnetizing inductance
Lr rotor inductance
Pn Rated power
Ppv PV power
Te Electromagnetic torque
Tr Load torque
Vdc DC-link voltage
Vdr , Vqr Rotor voltages
Vds , Vqs Stator voltages
ωm Motor speed
ωmn Nominal value of mechanical speed
ωsl Sliding speed
ϕdr Rotor flux
ϕrn Nominal value of flux

References
1. Ibrahim, M.N.; Rezk, H.; Al-Dhaifallah, M.; Sergeant, P. Solar Array Fed Synchronous Reluctance Motor
Driven Water Pump: An Improved Performance under Partial Shading Conditions. IEEE Access 2019, 7,
77100–77115. [CrossRef]
2. Ibrahim, M.N.; Rezk, H.; Al-Dahifallah, M.; Sergeant, P. Hybrid Photovoltaic-Thermoelectric Generator
Powered Synchronous Reluctance Motor for Pumping Applications. IEEE Access 2019, 7, 146979–146988.
[CrossRef]
3. Bouallègue, S.; Haggège, J.; Ayadi, M.; Benrejeb, M. PID-type fuzzy logic controller tuning based on particle
swarm optimization. Eng. Appl. Artif. Intell. 2012, 25, 484–493. [CrossRef]

33
Energies 2020, 13, 4473

4. Hu, H.; Hu, Q.; Lu, Z.; Xu, D. Optimal PID controller design in PMSM servo system via particle swarm
optimization. In Proceedings of the 31st Annual Conference of IEEE Industrial Electronics Society, IECON 2005,
Raleigh, India, 6–10 November 2005; pp. 79–83.
5. Hasanien, H.M.; Muyeen, S.M. Design optimization of controller parameters used in variable speed wind
energy Conversion system by genetic algorithms. IEEE Trans. Sustain. Energy 2012, 3, 200–208. [CrossRef]
6. Ambia, M.N.; Hasanien, H.M.; Al-Durra, A.; Muyeen, S.M. Harmony search algorithm-based controller
parameters optimization for a distributed-generation system. IEEE Trans. Power Deliv. 2015, 30, 246–255.
[CrossRef]
7. Blasko, V.; Kaura, V. A novel control to actively damp resonance in input LC filter of a three-phase voltage
source converter. IEEE Trans. Ind. Appl. 1997, 33, 542–550. [CrossRef]
8. Bingi, K.; Ibrahim, R.; Karsiti, M.N.; Chung, T.D.; Hassan, S.M. Optimal PID control of pH neutralization
plant. In Proceedings of the 2nd IEEE International Symposium on Robotics and Manufacturing Automation,
Ipoh, Malaysia, 25–27 September 2016; pp. 1–6.
9. Mishra, K.P.; Kumar, V.; Rana, K.P.S. Stiction combating intelligent controller tuning: A comparative study.
In Proceedings of the International Conference on Advances in Computer Engineering and Application,
Ghaziabad, India, 19–20 March 2015; pp. 534–541.
10. Ho, W.K.; Gan, O.P.; Tay, E.B.; Ang, E.L. Performance and gain and phase margins of well-known PID tuning
formulas. IEEE Trans. Control. Syst. Technol. 1996, 4, 473–477. [CrossRef]
11. Costa, B.L.G.; Graciola, C.L.; Angélico, B.A.; Goedtel, A.; Castoldi, M.F. Metaheuristics Optimization Applied
to PI controllers Tuning of a DTC-SVM Drive for Three-Phase Induction Motors. Appl. Soft Comput. 2017, 62,
776–788. [CrossRef]
12. Gizi, A.J.A.; Mustafa, M.; Jebur, H.H. A novel design of high-sensitive fuzzy pid controller. Appl. Soft Comput.
2014, 24, 794–805. [CrossRef]
13. Zemmit, A.; Messalti, S.; Harrag, A. A new improved dtc of doubly fed induction machine using ga based pi
controller. Ain Shams Eng. J. 2017, 4, 1–9. [CrossRef]
14. Premkumar, K.; Manikandam, B. Speed control of brushless dc motor using bat algorithm optimized adaptive
neuro-fuzzy inference system. Appl. Soft Comput. 2015, 32, 403–419. [CrossRef]
15. Mohamed, M.A.; Diab, A.A.; Rezk, H. Partial shading mitigation of PV systems via different meta-heuristic
techniques. Renew. Energy 2019, 130, 1159–1175. [CrossRef]
16. Abdalla, O.; Rezk, H.; Ahmed, E.M. Wind driven optimization algorithm based global MPPT for PV system
under non-uniform solar irradiance. Sol. Energy 2019, 180, 429–444. [CrossRef]
17. Oshaba, A.S.; Ali, E.S. Speed control of induction motor fed from wind turbine via particle swarm optimization
based pi controller. Res. J. Appl. Sci. Eng. Technol. 2013, 5, 4594–4606. [CrossRef]
18. Douiri, M.R.; Cherkaoui, M. Neuro-genetic observer speed for direct torque neuro-fuzzy control of induction
motor drive, Journal of circuits. Syst. Comput. 2012, 21, 1–18.
19. Ali, E.S. Speed control of induction motor supplied by wind turbine via imperialist competitive algorithm.
Energy 2015, 89, 593–600. [CrossRef]
20. Rajasekar, N.; Sundaram, K.M. Feed-back controller design for variable voltage speed induction motor drive
via ant colony optimization. Appl. Soft Comput. 2012, 12, 1566–1573. [CrossRef]
21. Sundareswaran, K.; Nayak, P.S. Ant Colony based feedback controller design for soft-starter fed induction
motor drive. Appl. Soft Comput. 2012, 12, 1566–1573. [CrossRef]
22. Salvatore, N.; Caponio, A.; Neri, F.; Stasi, S.; Cascella, G.L. Optimization of delayed state kalman filter based
algorithm via differential evolution for sensorless control of induction motors. IEEE Trans. Ind. Electron.
2010, 57, 385–394. [CrossRef]
23. Costa, B.L.G.; Angélico, B.A.; Goedtel, A.; Castoldi, M.F.; Graciola, C.L. Differential evolution applied to dtc
drive for three-phase induction motors using adaptive state observer. J. Control Autom. Electr. Syst. 2015, 26,
403–420. [CrossRef]
24. Hato, M.M.; Bouallègue, S.; Ayadi, M. Water Cycle Algorithm-tuned PI Control pf a Doubly Fed Induction
Generator for Wind Energy Conversion. In Proceedings of the 9th International Renewable Energy Congress
(IREC), Hammamet, Tunisia, 20–22 March 2018; pp. 1–6.
25. Hato, M.M.; Bouallègue, S. Direct Power Control Optimization for Doubly Fed Induction Generator Based
Wind Turbine Systems. Math. Comput. Appl. 2019, 24, 77.

34
Energies 2020, 13, 4473

26. Arfaoui, J.; Rezk, H.; Al-Dhaifallah, M.; Elyes, F.; Abdelkader, M. Numerical Performance Evaluation of
Solar Photovoltaic Water Pumping System under Partial Shading Condition using Modern Optimization.
Mathematics 2019, 7, 1123. [CrossRef]
27. Devanshu, A.; Singh, M.; Kumar, N. DSP based feedback linearization control of vector controlled induction
motor drive. In Proceedings of the 2016 IEEE 1st International Conference on Power Electronics, Intelligent
Control and Energy Systems (ICPEICES), Delhi, India, 4–6 July 2016; pp. 1–6.
28. Hamid, K.H.A.N. Field Oriented Control, Application Note; Polytechech: Clermont-Ferrand, France, 2008.
29. Devanshu, A.; Singh, M.; Kumar, N. Sliding Mode Control of Induction Motor Drive Based on Feedback
Linearization. IETE J. Res. 2020, 66, 256–269. [CrossRef]
30. Kim, S.-H. Maximum torque control of an induction machine in the field weakening region. IEEE Trans. Ind.
Appl. 1995, 31, 787–794.
31. Singh, B.; Shukla, S.; Chandra, A.; Al-Haddad, K. Loss minimization of two stage solar powered speed
sensorless vector controlled induction motor drive. In Proceedings of the IECON 2016—42nd Annual
Conference of the IEEE Industrial Electronics Society, Florence, Italy, 24–27 October 2016; pp. 1–6.
32. Salim, D.; Kheldoun, A.; Sadouni, R. Fuzzy indirect field oriented control of dual star induction motor water
pumping system fed by photovoltaic generator. Int. J. Eng. Intell. Syst. Electr. Eng. Commun. 2015, 23, 63–76.
33. Singh, S.; Singh, B. Solar PV water pumping system with DC-Link voltage regulation. Int. J. Power Electr.
2015, 7, 72–85. [CrossRef]
34. Wanzeller, A.M. Current control loop for tracking of maximum power point supplied for photovoltaic array.
IEEE Trans. Instrum. Meas. 2004, 53, 1304–1310. [CrossRef]
35. Marouani, R.; Mami, A. Voltage Oriented Control Applied to a Grid Connected Photovoltaic System with
Maximum Power Point Tracking Technique. Am. J. Appl. Sci. 2010, 7, 1168–1173. [CrossRef]
36. Consoli, A.R. Experimental low-chattering sliding-mode control of a pm motor drive. Eur. Power Electr. 1991,
1, 13–18.
37. Fathy, A.; Al-Dhaifallah, M.; Rezk, H. Recent Coyote Algorithm-Based Energy Management Strategy for
Enhancing Fuel Economy of Hybrid FC/Battery/SC System. IEEE Access 2019, 7, 179409–179419. [CrossRef]
38. Pierezan, J.; Coelho, L.D.S. Coyote optimization Algorithm: A new metaheuristic for global optimization
problems. In Proceedings of the 2018 IEEE Congress on Evolutionary Computation (CEC), Rio de Janeiro,
Brazil, 29 January 2018; pp. 1–8.

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

35
energies
Article
Performance Analysis of Axial-Flux Induction Motor
with Skewed Rotor
Fatma Keskin Arabul *, Ibrahim Senol and Yasemin Oner
Department of Electrical Engineering, Yildiz Technical University, 34220 Istanbul, Turkey;
[email protected] (I.S.); [email protected] (Y.O.)
* Correspondence: [email protected]

Received: 27 August 2020; Accepted: 21 September 2020; Published: 23 September 2020

Abstract: In recent years, with developing technology in the field of electrical machines, more efficient
and high power density electric motors have been produced. The use of high energy efficiency motors
gains importance due to the increase in global energy demand. The main purpose of this study was
to design an Axial Flux Induction Motor (AFIM) with the same efficiency class as the Radial Flux
Induction Motor (RFIM) in premium efficiency (IE3) class which is used commonly in industrial
applications. Various AFIMs are designed with different rotor slot numbers and performance analyses
as efficiency and torque ripple changes are investigated. It is known that torque ripple is one of
the key parameters in electrical machine design which should be kept as low as possible without
decreasing efficiency and torque. Accordingly, AFIMs’ rotor slots are skewed considering the stator
and rotor slot numbers. The use of a Soft Magnetic Composites (SMC) material in design is also
investigated. As a result of the analyses, many premium efficiency classes for AFIMs are obtained.
In addition, using SMC material and skewing the rotor slots provides that torque ripples be reduced.

Keywords: axial flux induction motor; finite element analysis; performance evaluation

1. Introduction
Today, more than 40% of the global energy consumption amount is consumed by induction motors
and this rate exceeds 70% in the industry [1,2]. Additionally, these motors are key components of
many industrial processes, with their reliability, and low cost of maintenance and construction [3,4].
Considering the amount of energy consumed by these motors, it is seen that even a small change in
their efficiency will provide significant savings in worldwide energy consumption. The largest energy
savings, particularly for medium and small power motors, arise for their higher efficiency classes [5,6].
Accordingly, the use of IE3 class motors has become mandatory due to the laws published in many
countries. For instance, IE3 motors have been mandatory since 2011 in the United States and in Turkey,
China, and the EU countries in 2015 [7–9]. Nowadays, energy consumption and environmental impacts
are reduced with high efficiency motors. Additionally, motor reliability increases sustainable use and
demand for investment [10]. In many countries around the world, many programs are encouraged
to increase the use of high-efficiency motors. Among these programs, the Efficiency Increasing
Project and the Efficient Motor Replacement programs are prominent in the world. Replacing existing
low-efficiency motors with high-efficiency motors will result in significant energy savings even if
same-sized motors are used [11,12]. Otherwise, if AFIMs are preferred over conventional RFIMs,
more efficient and smaller volume motors can be designed [13,14].
With the developing technology, electric motors have a more compact structure and many studies
are carried out to increase their efficiency. In many studies, instead of radial flux design of induction
motors, axial flux design was found to have a more compact structure and it was concluded that their
efficiency and torque density could be increased further [14,15].

Energies 2020, 13, 4991; doi:10.3390/en13194991 www.mdpi.com/journal/energies

37
Energies 2020, 13, 4991

AFIMs have the same operating principle as RFIM. However, design of these motors is quite
different. The main difference is the magnetic flux direction. In conventional radial flux machines,
the flux is in radial direction relative to the machine axis. The magnetic flux produced in AFIMs is in
axial direction with respect to the machine axis.
In recent decades, AFIMs have been a popular research topic for researchers. Many studies have
been conducted in the literature on AFIM design and control [16–18]. Among these studies, the most
notable ones are the design and implementation of different structures. Additionally, many AFIMs
have been designed for various applications such as pumps, electric vehicles, wind turbine, etc. [19,20]
For instance, in electric vehicles, which are among the popular topics of today, AFIMs have different
uses such as wheel-directly coupled, on-wheel, or main motor [21]. In some studies, different materials
have been used, such as iron, magneto dielectric, superconductors, etc. [22,23].
In this study, AFIMs are designed to have the same efficiency class as RFIM in IE3 efficiency class
used in industry. In addition, the rotor is skewed to minimize the torque ripple. In the literature,
only stator slot numbers has been taken into consideration to determine the skew angles for the
performance analysis of an induction motor [24–27]. In this study, and different from the literature,
skew angles are selected considering both the ratio of stator and rotor slot numbers. Additionally,
analyses are carried out using SMC material for stator, rotor, and both stator and rotor of the motor,
which provides the best results in terms of efficiency and torque ripple. However, in the analysis
results obtained, a considerable decrease in torque ripple over 11% is observed.
Following the introduction, Section 2 presents the topologies of AFMs and design parameters
of AFIM. Section 3 introduces model properties, results, and discussion of all various types of AFIM
models. Finally, the concluding remarks are presented in Section 4.

2. Methodology
In this section, considering the geometrical properties of the stator/rotor core, radial and axial flux
machine equations are presented after the topologies of AFMs are introduced.

2.1. Axial Flux Machine Topologies


A machine with one air gap is the oldest and simplest structure of AFM, has a single-sided
motor, and a single stator-single rotor (SSSR). The structure of this machine is easier due to a single air
gap [28]. Generally, this machine is the best choice for low torque applications such as fans, pumps,
food processors, etc. [29] Also, it can be said that single-sided AFIMs are more resistant to static
eccentricity than conventional motors [30,31]. The disadvantage of this type of machine is that bearing
life depends on their load. Active material utilization of the SSSR machine is higher [32]. In Figure 1,
single-sided AFIM components are shown.

Figure 1. Single-sided Axial Flux Induction Motor (AFIM) components.

38
Energies 2020, 13, 4991

There are two air gaps in this type of machine. Both of these air gaps can be axially (double-sided
motor), or one axially and the other radially. Such motors consist of a double stator-single rotor,
with the rotor sandwiched between the stators, or single stator-double rotor with the stator sandwiched
between the rotors. The advantages of double-sided motors include high torque density and balanced
axial forces.
In this context, and in terms of economy, the production of two stators is more costly, especially in
small powerful machines compared to the single-sided structure. However, the difference in production
cost between single-sided motor and double-sided motor for high torque machines is decreasing.
In double-sided motors, the moment of inertia is lower and the rotor is lighter [33].
Although the double-sided motor structure has better performance, for high-powered motors,
the multi-air-gap disc structure is a better choice [34]. Multi air gap machines have two topologies
that are determined by the number of stators and rotors. If the number of stators is more than the
number of rotors, these machines are called external stator and internal rotor machines. If the number
of rotors is more than the number of stators, these machines are called internal stator external rotor type
machines. Internal stator external rotor type machines are preferred due to their high efficiency [35].
This topology can be defined as a concept rather than a machine type. The aim is to place the stators
and rotors alternately to meet the requirements of the application. An advantage of this configuration
is that it offers modularity [36].
In this study, single air-gap motor topology—which is also prominent in terms of ease of design
and analysis—is chosen. It is an advantage that the volume of this structure is smaller. A conventional
radial flux induction motor used industrially in the premium efficiency class is taken as a reference to
the designed AFIMs.

2.2. Design Parameters of AFIM


The rotating magnetic field can be solved analytically by integrating the basic flux with respect to
all radius (r) and pole form factor (α). The rotating magnetic field is within a pole range and electrical
angle values are accepted. Axial rotating magnetic field is obtained as in Equation (1).

2Bmax 2 2
φax = (r2 − r1 ) cos ωt (1)
p

where φax is axial rotating magnetic field in disk-like air gap; Bmax is maximum induction in the air
gap; p is number of poles; r2 is outer radius of the core of axial induction motor and r1 is inner radius
of the core of axial induction motor.
Basically, the output power of electrical machines (P2 ) is a function of flux per pole as in
Equation (2).
P2 = f (φ) (2)

Therefore, both types of induction motors can be compared using the flux equations of the rotating
magnetic field. In 1986, Varga compared both types of motors using the flux equations of the rotating
magnetic field in his study [37]. Thus, the flux equation of AFIM (Equation (1)) has been compared
with the similar equation of RFIM (Equation (3)).

4Bmax
φrad = Lr (rr − ro ) cos ωt (3)
p

where φrad is radial rotating magnetic field in cylindrical air gap; Lr is length of the radial induction
motor, rr is rotor radius of the radial induction motor; ro is radius of shaft opening.
Flux values for two types of motors are equalized for predictive comparison (φax = φrad ),
so Equation (4) becomes;
2Lr (rr − ro ) = r22 − r21 (4)

39
Energies 2020, 13, 4991

In fact, both sides of equation 4 are cross-sectional areas for total magnetic flux in related type
induction motors. In Figure 2, RFIM and AFIM geometries are shown. This comparison takes into
account only the geometric properties of different types of induction motors under equal magnetic use.

Figure 2. (a) Radial Flux Induction Motor (RFIM) geometry, (b) AFIM geometry.

The shaft opening increases the rotor radius rr by ro . The equations given AFIM have a single
air gap and no volumetric assumptions are made. In Equation (5), apparent internal power is solved;
r1 and r2 are halves of D1 and D2 , respectively.

Si = Cax (D22 − D21 )n1 (5)

where Si is apparent internal power; Cax is axial induction motor constant; D2 is core outside diameter
with no slots; D1 is core inside diameter; n1 is synchronous speed.
In Equation (6), Cax is calculated as;

π2 αkw Bmax F1
Cax = √ Bmax F1 = (6)
2480 109.13

where α is deflection, pole form factor; kw is winding factor; and F1 is distributed MMF; kw and α
values are taken 0.9 and 0.7, respectively, as the average values usually used [38].
In Equation (7), Da refers to the average diameter of the core;

D2 + D1
Da = (7)
2
In Equation (8), L refers to the radial width of the core;

D2 − D1
L= (8)
2
While establishing equations for diameter dimensions of AFIM, slots are neglected. The motor air
gap is assumed to be constant when the slots are neglected. In these conditions, the outer diameter of
the core D2 can be found as in Equation (9).

2pφax
D2 = D21 + (9)
Bmax

40
Energies 2020, 13, 4991

Core volume of the entire stator and rotor (V) is calculated in Equation (10).
π 2 
V = Vs + Vr = (D − D21 ) hi (10)
4 2
i

where Vs is core volume of stator; Vr is core volume of rotor; hi is total height of the machine.
i
Efficiency is calculated as in Equation (11) according to IEEE Standard 112-2017 [39].

Poutput Pinput − Plosses
η= = (11)
Pinput Pinput

where η is efficiency in percent; Poutput is output power; Pinput is inner output [39]. Plosses is total
losses and presented as in Equation (12);

Plosses = Pcore + Pcu + PFW + PSTR (12)

where Pcore is stator and rotor core losses; Pcu is stator and rotor copper losses; PFW is mechanical
losses as frictional and windage losses and PSTR is stray-load losses. The biggest cause of stray-load
losses is harmonic energies when the motor is operating under load. If the load current consists of
harmonic components, flux magnitude and waveform were distorted, which results in harmonic
torques, vibrations and the noise [40,41]. In the IEEE Standard 112-2017, the assumed values for
stray-load losses are shown in Table 1. According to the table, in the study, stray-load losses percent of
rated load are taken as 1.8% for designed AFIMs [39].

Table 1. Assumed values for stray-load loss.

Machine Rating (kW) Stray-Load Loss Percent of Rated Load


0.7457 to 90 1.8%
91 to 375 1.5%
376 to 1850 1.2%
1851 and greater 0.9%

In this study, transient analyses are carried out and torque values are calculated with the help of
ANSYS Maxwell program. Torque ripple (τripple ) in percent is defined as the ratio of difference between
maximum (τmax ) and minimum torque (τmin ) values to average torque (τavg ) as in Equation (13).

τmax − τmin
τripple = · 100 (13)
τavg

AFIM results based on different rotor slot numbers are examined in the next section. In this study,
the rotor slots are skewed to achieve better efficiency and fewer torque ripple. Skew angles are given
for all slot numbers of the non-skew models, and analyses have been made by taking into account the
stator and rotor slot numbers. Skew angle analyses adjusted according to stator slot number are given
in Equations (14)–(16).
360
βs = (14)
Ns
360 ◦
1.5 · βs = = 15 (15)
Ns
360 ◦
2 · βs = = 20 (16)
Ns

41
Energies 2020, 13, 4991

where βs is skew angle according to stator slot number in degrees, Ns is stator slot number. By using
Equations (14)–(16); 10◦ , 15◦ and 20◦ are selected. These values are the same for all design and analyzed
for all of them.
Skew angle analyses adjusted according to rotor slot number are given in Equations (17) and (18).

360
βr = (17)
Nr

360
2 · βr = (18)
Nr
where βs is skew angle according to rotor slot number in degrees, Nr is rotor slot number.
In this context, geometries with three additional skew angles, βs , 1.5 · βs , 2 · βs , βr and 2 · βr are
applied in the rotor slots of the 28, 30, 32, 34, 38, 40 and 42 slotted models, and analyses results
are presented.

3. Results and Discussion


In this section, general design parameters of analyzed AFIMs with reference RFIM are presented.
In addition, many AFIMs are designed to be in the same energy efficiency class as the RFIM referenced.
While making these designs, no changes are made on the stator. When using the same stator, rotor slot
numbers and skew angles are changed.

3.1. RFIM Model Details


This section provides information about the reference RFIM, which is widely used in the industrial
applications. The analysis results of this motor, which is IE3 efficiency class with 93.2% efficiency and
7.49% torque ripple, was made by using the ANSYS Maxwell program. Analyses in which the motor
model is suitable for 2D analysis are done in two dimensions. The ANSYS model of the model is
shown in Figure 3. In addition, general design parameters of RFIM are shown in Table 2.

 
(a) (b)

Figure 3. (a) RFIM model (b) Magnetic field density of RFIM.

42
Energies 2020, 13, 4991

Table 2. Design parameters of Radial Flux Induction Motor (RFIM).

Parameter Value
Rated power 30 kW
Rated voltage 400 V
Rated speed 2980 rpm
Rated frequency 150 Hz
Stator core outer diameter 355 mm
Rotor core outer diameter 236.5 mm
Length of motor 250 mm
Stator/Rotor slot numbers 45/40
Stator material Steel (M19_24G)

3.2. AFIM Model Details


In this section, Finite Element Analysis (FEA) is needed to perform electromagnetic analysis of
AFIM. Unlike other types of machines, 3D analysis is mandatory because 2D analysis of such machines
does not give detailed results. For this reason, analyses are performed in 3D Cartesian coordinate
system and these analyses take a considerable amount of time due to high meshing. Also, for detailed
analysis, the number of meshes is kept as high as the processor allowed. In this study, ANSYS Maxwell
software is preferred to the FEA analysis, since it is suitable for magnetic field analysis of machines
and has a wide usage area.
Firstly, an AFIM designed with the same number of stator and rotor slots. AFIM model and the
magnetic field density of AFIM is shown in Figure 4. This model is designed with a non-skewed
rotor. According to the analysis results, the efficiency of AFIM is 82.83%, while the torque ripple
is 7.4%. Thereafter, in the next section, AFIMs are designed with different rotor slot numbers and
various skewed rotors to increase the efficiency to IE3 efficiency class. Table 3 shows the general design
parameters of the analyzed AFIMs.

Table 3. General design parameters of Axial Flux Induction Motor (AFIMs).

Parameter Value
Rated power 30 kW
Rated voltage 400 V
Rated speed 2895 rpm
Rated frequency 150 Hz
Stator/Rotor core outer diameter 355 mm
Stator/Rotor core inner diameter 140 mm
Air-gap length 2 mm
Length of stator/rotor 60/55 mm
Stator slot numbers 36
Rotor slot numbers 28, 30, 32, 34, 38, 40, 42
Stator material Steel (M19_24G)
Winding type Whole Coiled
Number of conductors per slot 16
Number of strands 2
Conductor type Copper

43
Energies 2020, 13, 4991

 
(a) (b)

Figure 4. (a) AFIM model (b) Magnetic field density of AFIM.

3.3. AFIM Analyze Results with Different Rotor Slot Numbers


In this section, analyses are made according to seven different slot numbers for the rotor slots
without skewing, 28, 30, 32, 34, 38, 40 and 42 slots, respectively. While determining these numbers,
rotor slot widths are taken into account according to the machine size. Therefore, the maximum slot
number is specified as 42 and the minimum as 28 for the number of rotor slots. Since the stator slot
number is 36, 36-slot rotor design is not feasible, and because the stator slot and rotor slot areas are
overlapped, the magnetic flux production is not possible.
It is known that torque ripple is one of the most challenging parameters in electrical machine
designs. While making designs, decreasing this value is an aim. Therefore, torque ripples are taken
into consideration while making comparisons. Figure 5 shows the efficiency and torque ripples of the
analysis results according to 7 different slot numbers from 28 to 42. In Figure 5, torque ripple in 30 and
42 slotted AFIMs varies a lot compared to others. As the reason for this, it can be said that the number
of these two slots is a multiple of 6; that is, the pole number. Thus, they are forced more in magnetic
field transition. From the other five AFIMs, the torque ripples of only 28 and 34 slotted rotor designs
are over 10%, while the other three designs’ torque ripples at 9%. In this way, AFIM models with the
best performance in terms of torque ripples are 32, 38 and 40 slotted models. There are six models in
the same energy efficiency class (IE3) as RFIM. Only 34 slotted model is in a lower efficiency class (IE2).

Figure 5. Efficiency and torque results of AFIM with different slot numbers.

As a result of the analyses made according to different rotor slot numbers, although it does
not have the highest efficiency, it can be said that the 40 slotted model, which is in the same energy
efficiency class and has little difference between its efficiency, is the best model among them since the
torque ripple is much less.

44
Energies 2020, 13, 4991

3.4. Results of AFIM with Skewed Rotor


In the analysis of the rotor with 28 slot non-skew model, torque ripple 13.9%, efficiency 92.6% are
obtained. The efficiency class of this model is in IE3 class. As the first analysis, skew angles are given to
28-slot AFIM. Accordingly, the skew angles in analyses are 10◦ , 12.86◦ , 15◦ and 20◦ . Torque ripple and
efficiency results of the analyses are presented in Figure 6. Modeling of 25.72◦ skew is not feasible due
to the overlay occurring between stator and rotor. In Figure 6, it is observed that there is no increase in
efficiency when the rotor is skewed. However, when the models with skewed according to the number
of stator slots are examined, it is seen that the efficiency decreases as the skew angle increases. On the
other hand, when the skew is given according to the number of rotor slots, the efficiency keeps the
same value.

Figure 6. Efficiency and torque results of AFIM with 28 slots.

As a result of the non-skew model analyses, 30-slot AFIM model has the highest torque ripple
value and the best efficiency. In the second analysis, this model is skewed. In addition to the non-skew
model; in Figure 7 the torque ripple and the efficiency results of 10◦ , 12◦ , 15◦ and 20◦ skewed AFIM
models are presented. In Figure 7, the torque ripple has decreased as the skew angle degree increases
in analyses made only considering the number of stator slots. When analyses made according to the
number of rotor slots are compared among themselves, it is seen that as the skew angle increases,
the torque ripple decreases. In Figure 7, an increment in efficiency was not observed when the rotor is
skewed. However, according to the skewed models, efficiency is increasing with the increment in the
skew angle.

Figure 7. Efficiency and torque results of AFIM with 30 slots.

45
Energies 2020, 13, 4991

Considering the number of stators and rotor slots, the third analyses are done to 32-slot AFIM
model with the skew angle of 10◦ , 11.25◦ , 15◦ , 20◦ and 22.5◦ . The non-skewed 32-slot AFIM model’s
torque ripple is 9.5% and the efficiency is 92.2%, which is in the IE3 class. Torque ripple and efficiency
results of the analyses are shown in Figure 8. A comparison of the non-skew and 10◦ skew models
shows that 10◦ skewed model has less torque ripple. On the other hand, a comparison of only skewed
models shows that the torque ripple is lowest at 10◦ and highest at 15◦ . Comparing the 32-slot skewed
models and the non-skew model, it is seen that the torque ripple decreases when the skew angle is 10◦ .

Figure 8. Efficiency and torque results of AFIM with 32 slots.

In addition, efficiency increases and the efficiency class increases from IE2 to IE3. Only from
skewed models according to the number of stator slots, the efficiency decreases while the skew angle
increases. On the other hand, according to the skewed models skewed with rotor slots number,
the torque ripple increases and the efficiency decreases while the skew angle increases.
The fourth analysis is performed to 34-slot AFIM and the skew angles are calculated as 10◦ , 10.58◦ ,
15 , 20◦ and 21.17◦ . In the 34-slot non-skewed model, the torque ripple is 11%, the efficiency is 90.8%,

while the efficiency class is in the IE2 class. Torque ripple and efficiency results of the 32-slot AFIM
analyses are shown in Figure 9. It is observed that the torque ripple does not fluctuate very much.
Accordingly, it can be seen that the 15◦ skewed model is the most efficient model.

Figure 9. Efficiency and torque results of AFIM with 34 slots.

In the 5th analysis step, the 38-slot AFIM, which has the least torque ripple, is analyzed by
considering the stator and rotor slots numbers. Accordingly, skew angles are determined as 9.47◦ ,
10◦ , 15◦ , 18.95◦ and 20◦ . When the skewed analysis results are compared to the non-skewed model,

46
Energies 2020, 13, 4991

the analysis with the lowest torque ripple is obtained when the skew angle is 9.47. Torque ripple and
efficiency results of the 38-slot AFIM analyses are shown in Figure 10.

Figure 10. Efficiency and torque results of AFIM with 38 slots.

When the results of the skewed models of the 38-slot AFIM model according to the number of
stator and rotor slots are examined, we do not see a positive effect on efficiency. Although the results
are close to each other, the non-skewed model gives better results than skewed models.
In the sixth analysis step, the skew applied to the rotor of the 40-slot AFIM, which has the second
best efficiency among non-skewed models. This model is also in IE3 energy efficiency class and the
only model with low torque fluctuation. Skew angles for this model are calculated as 9◦ , 10◦ , 15◦ , 18◦
and 20◦ . Figure 11 shows the torque and efficiency results of the non-skewed and skewed models.
In Figure 11, the 18◦ skewed model has the best torque ripple percentage. It can be seen that giving a
skew to the rotor of 40-slot AFIM affects the efficiency negatively. However, the non-skewed, 9◦ , 10◦ ,
15◦ and 18◦ skewed models are in IE3 energy efficiency class, other 10◦ and 20◦ skewed models are in
IE2 energy efficiency class.

Figure 11. Efficiency and torque results of AFIM with 40 slots.

The last step of analyses among skewed models is done to 42-slot AFIM, which is another model
with high torque ripple. Analyses results of non-skewed and skewed models of 42-slot AFIM are
shown in Figure 12. The non-skewed model of 42-slot AFIM is in IE3 energy efficiency class. According
to the analyses results, it is observed that there is a different situation between the efficiencies that are
not encountered in the other slot numbers. Efficiency decreased in models with 10◦ and 20◦ skewed,
whereas efficiency in skew angle 8.57◦ and 15◦ is significantly increased. In fact, it has been observed

47
Energies 2020, 13, 4991

that three of these models are in the IE4 energy efficiency class. These are the highest efficient models
achieved among analyses.

Figure 12. Efficiency and torque results of AFIM with 42 slots.

As a result of all analyses carried out, giving skews to the rotor of the motor affects the efficiency
differently, and the torque ripple in different rotor slot numbers of AFIMs. In some AFIMs, skewing
has a positive effect in terms of torque ripple and decreases the torque ripple, while in others it is
observed that the torque ripple increases. In the same way, it has different effects in terms of efficiency.
For example, the 8.57◦ , 15◦ and 17.14◦ skewed 42-slot AFIMs are given results in IE4 energy efficiency
class. Also, the torque ripple of the same model decreases by 28% to 26.6%. Therefore, the best
configuration among 42-slot models is obtained as 8.57◦ skew angle.
However, in the analyses performed both in terms of torque ripple and efficiency, the most
optimum condition result is achieved in the design with a 10◦ skew given to the 32-slot AFIM.
The torque ripple of this model is 8.93%, the effıciency is 93.1% and efficiency class of this motor is IE3.
In the next section, the effect of the material on efficiency and torque ripple by using SMC material
is examined.

3.5. Results of AFIM with SMC Material


In this section, the effect of the SMC material on efficiency and torque is examined. By this
purpose, three analyses are done with using SMC material for stator, rotor and both of them. Somaloy
700-3P is selected as a SMC material with a density 7.57 g/cm3 [42].
According to the analyses results, by changing the material, the efficiencies stayed in the same
efficiency class as IE3. However, only SMC stator has the best efficiency value, both SMC stator&rotor
has the best torque value. Compared to the non-skewed based model to both SMC stator&rotor used
model over 11% of reduction is observed in torque ripple. Additionally, changing the material has a
positive impact on the efficiency. Figure 13 shows the efficiency and torque results of SMC used AFIM
by different parts.
In the X-axis of Figure 13, skew status and material of AFIM are defined as follows:

(a) Rotor is not skewed and both stator and rotor material are steel,
(b) Rotor is skewed and both stator and rotor material are steel,
(c) Rotor is skewed and stator material is SMC and rotor material is steel,
(d) Rotor is skewed and stator material is steel and rotor material is SMC,
(e) Rotor is skewed and both stator and rotor material are SMC.

48
Energies 2020, 13, 4991

Figure 13. Efficiency and torque results of SMC used AFIM.

4. Conclusions
Since the aim in this study is to achieve an AFIM in the same energy efficiency class as RFIM,
the efficiency of AFIMs with torque ripples are investigated. In this study, transient analyses are carried
out and torque values are calculated with the help of ANSYS Maxwell program. Efficiency values
of analysis results according to seven different slot numbers from 28 to 42 are presented. In addition
to the efficiency values, the percentage torque ripples and energy efficiency class are also examined.
To reduce the torque ripple values, rotor is skewed according to stator and rotor slot numbers. It was
seen that the effects of skewing on efficiency and torque ripple are variable. Also, it was observed
that to choose the skewing angle degrees, not only with stator slot number, but also rotor slot number
could affect positive the performance of AFIMs. Using a SMC material on different parts of AFIM had
a positive impact on performance analysis of a model, as reducing the torque ripple over 11% while
the models are in premium efficiency class. For further studies, different materials could be used for
various applications.

Author Contributions: F.K.A., I.S., Y.O. work on conceptualization, methodology, software, validation,
formal analysis, writing—original draft preparation, writing—review and editing. All authors have read
and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Acknowledgments: The authors declare no acknowledgements.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. De Almeida, A.T.; Ferreira, F.J.T.E.; Fong, J.A.C. Standards for Efficiency of Electric Motors. IEEE Ind.
Appl. Mag. 2010, 17, 12–19. [CrossRef]
2. Ferreira, F.J.T.E.; Leprettre, B.; De Almeida, A.T. Comparison of Protection Requirements in IE2-, IE3-,
and IE4-Class Motors. IEEE Trans. Ind. Appl. 2016, 52, 3603–3610. [CrossRef]
3. Elvira-Ortiz, D.A.; Morinigo-Sotelo, D.; Zorita-Lamadrid, A.L.; Osornio-Rios, R.A.; Romero-Troncoso, R.d.J.
Fundamental frequency suppression for the detection of broken bar in induction motors at low slip and
frequency. Appl. Sci. 2020, 10, 4160. [CrossRef]
4. Pietrowski, W.; Górny, K. Analysis of Torque Ripples of an Induction Motor Taking into Account a Inter-Turn
Short-Circuit in a Stator Winding. Energies 2020, 13, 3626. [CrossRef]
5. Esen, G.K.; Ozdemir, E. A New Field Test Method for Determining Energy Efficiency of Induction Motor.
IEEE Trans. Instrum. Meas. 2017, 66, 3170–3179. [CrossRef]
6. Parviainen, A. Design of Axial-Flux Permanent-Magnet Low-Speed Machines and Performance Comparison between
Radial-Flux and Axial-Flux Machines; Lappenranta University of Technology: Lappeenranta, Finland, 2005;
ISBN 9522140295. Available online: https://lutpub.lut.fi/bitstream/handle/10024/31185/TMP.objres.74.pdf?
sequence=1 (accessed on 22 September 2020).

49
Energies 2020, 13, 4991

7. Boglietti, A.; Cavagnino, A.; Vaschetto, S. Induction motor EU standards for efficiency evaluation: The scenario
after IEC 60034-2-1. In Proceedings of the IECON 2011—37th Annual Conference of the IEEE Industrial
Electronics Society, Melbourne, VIC, Australia, 7–10 November 2011; pp. 2786–2791.
8. IEC-Governments & International Organizations, Examples by Industry Sector: Electric Motors-Measuring
Efficiency. Available online: https://www.iec.ch/perspectives/government/sectors/electric_motors.htm
(accessed on 18 March 2020).
9. Electric Motors | European Commission. Available online: https://ec.europa.eu/info/energy-climate-change-
environment/standards-tools-and-labels/products-labelling-rules-and-requirements/energy-label-and-
ecodesign/energy-efficient-products/electric-motors_en (accessed on 18 March 2020).
10. De Almeida, A.T.; Ferreira, F.J.T.E.; Baoming, G. Beyond induction motors—Technology trends to move up
efficiency. IEEE Trans. Ind. Appl. 2014, 50, 2103–2114. [CrossRef]
11. Ferreira, F.J.T.E.; Baoming, G.; De Almeida, A.T. Reliability and operation of high-efficiency induction motors.
In Proceedings of the 2015 IEEE/IAS 51st Industrial and Commercial Power Systems Technical Conference, I and CPS
2015, Calgary, AB, Canada, 5–8 May 2015; Institute of Electrical and Electronics Engineers Inc.: New York, NY,
USA, 2015.
12. The Impact of New IE3 Requirements on Motors and Controls. Available online: https://www.
totallyintegratedautomation.com/2017/11/impact-new-ie3-requirements-motors-controls/ (accessed on
18 March 2020).
13. Valtonen, M.; Parviainen, A.; Pyrhönen, J. Electromagnetic field analysis of 3D structure of axial-flux solid-rotor
induction motor. In Proceedings of the International Symposium on Power Electronics, Electrical Drives,
Automation and Motion, Taormina, Italy, 23–26 May 2006; SPEEDAM 2006. Volume 2006, pp. 174–178.
14. Evans, P.D.; Eastham, J.F. Disc-geometry homopolar synchronous machine. IEE Proc. B Electr. Power Appl.
1980, 127, 299–307. [CrossRef]
15. Valtonen, M.; Parviainen, A.; Pyrhönen, J. The effects of the number of rotor slots on the performance
characteristics of axial-flux aluminium-cage solid-rotor core induction motor. In Proceedings of the IEEE
International Electric Machines & Drives Conference, Antalya, Turkey, 3–5 May 2007; Volume 1, pp. 668–672.
16. Dianati, B.; Kahourzade, S.; Mahmoudi, A. Analytical design of axial-flux induction motors. In Proceedings of
the 2019 IEEE Vehicle Power and Propulsion Conference, VPPC 2019—Proceedings, Hanoi, Vietnam, 14–17 October
2019; Institute of Electrical and Electronics Engineers Inc.: New York, NY, USA, 2019.
17. Rodger, D.; Coles, P.C.; Allen, N.; Lai, H.C.; Leonard, P.J.; Roberts, P. 3D finite element model of a disc
induction machine. In Proceedings of the IEE Conference Publication, Cambridge, UK, 1–3 September 1997;
pp. 148–149.
18. Valtonen, M.; Parviainen, A.; Pyrhönen, J. Inverter switching frequency effects on the rotor losses of
an axial-flux solid-rotor core induction motor. In Proceedings of the POWERENG 2007—International
Conference on Power Engineering—Energy and Electrical Drives Proceedings, Setubal, Portugal,
12–14 April 2007.
19. Dianati, B.; Kahourzade, S.; Mahmoudi, A. Optimization of Axial-Flux Induction Motors for the Application
of Electric Vehicles Considering Driving Cycles. IEEE Trans. Energy Convers. 2020, 35, 1522–1533. [CrossRef]
20. Mei, J.; Lee, C.H.T.; Kirtley, J.L. Design of axial flux induction motor with reduced back iron for electric
vehicles. IEEE Trans. Veh. Technol. 2020, 69, 293–301. [CrossRef]
21. Benoudjit, A.; Guettafi, A.; Nait Saïd, N. Axial flux induction motor for on-wheel drive propulsion system.
Electr. Mach. Power Syst. 2000, 28, 1107–1125. [CrossRef]
22. Álvarez, A.; Suárez, P.; Cáceres, D.; Cordero, E.; Ceballos, J.M.; Pérez, B. Disk-shaped superconducting rotor
under a rotating magnetic field: Speed dependence. IEEE Trans. Appl. Supercond. 2005, 15, 2174–2177.
[CrossRef]
23. Kubzdela, S.; Weglinski, B. Magnetodielectrics in Induction Motors with Disk Rotor. IEEE Trans. Magn. 1987,
24, 635–638. [CrossRef]
24. Beleiu, H.G.; Maier, V.; Pavel, S.G.; Birou, I.; Pica, C.S.; Darab, P.C. Harmonics consequences on drive systems
with induction motor. Appl. Sci. 2020, 10, 1528. [CrossRef]
25. Carbonieri, M.; Bianchi, N. Induction motor rotor losses analysis methods using finite element. In Proceedings
of the 2020 IEEE International Conference on Industrial Technology (ICIT), Buenos Aires, Argentina,
26–28 February 2020; pp. 187–192.

50
Energies 2020, 13, 4991

26. Nobahari, A.; Darabi, A.; Hassannia, A. Various skewing arrangements and relative position of dual rotor
of an axial flux induction motor, modelling and performance evaluation. IET Electr. Power Appl. 2018, 12,
575–580. [CrossRef]
27. Le Besnerais, J.; Lanfranchi, V.; Hecquet, M.; Romary, R.; Brochet, P. Optimal slot opening width for magnetic
noise reduction in induction motors. IEEE Trans. Magn. 2009, 45, 3131–3136. [CrossRef]
28. Hecker, Q.; Igelspacher, J.; Herzog, H.G. Parameter identification of an axial-flux induction machine by
winding functions. In Proceedings of the 19th International Conference on Electrical Machines- ICEM 2010,
Rome, Italy, 6–8 September 2010.
29. Chan, C.C. Axial-field electrical machines—Design and applications. IEEE Trans. Energy Convers. 1987.
[CrossRef]
30. Nasiri-Gheidari, Z.; Lesani, H. New design solution for static eccentricity in single stator-single rotor axial
flux induction motors. IET Electr. Power Appl. 2013, 7, 523–534. [CrossRef]
31. Nobahari, A.; Darabi, A.; Hassannia, A. Axial flux induction motor, design and evaluation of steady
state modeling using equivalent circuit. In Proceedings of the 8th Power Electronics, Drive Systems and
Technologies Conference, PEDSTC 2017, Mashhad, Iran, 14–16 February 2017.
32. Patterson, D.J.; Colton, J.L.; Mularcik, B.; Kennedy, B.J.; Camilleri, S.; Rohoza, R. A comparison of radial and
axial flux structures in electrical machines. In Proceedings of the 2009 IEEE International Electric Machines
and Drives Conference, IEMDC ’09, Miami, FL, USA, 3–6 May 2009; pp. 1029–1035.
33. Mirzaei, M.; Mirsalim, M.; Abdollahi, S.E. Analytical modeling of axial air gap solid rotor induction machines
using a quasi-three-dimensional method. IEEE Trans. Magn. 2007, 43, 3237–3242. [CrossRef]
34. Nasiri-Gheidari, Z.; Lesani, H. A survey on axial flux induction motors. Prz. Elektrotechniczny 2012, 2,
300–305.
35. Babu, V.R.; Soni, M.P. Modelling of Twin Rotor Axial Flux Induction Machine and its Application as
Differential in Electrical Vehicles. IJIREEICE 2017, 5, 118–127. [CrossRef]
36. Egea, A. Overview of Axial Flux Machines. Electr. Energy Mag. 2013, 4, 2–13.
37. Varga, J.S. Magnetic and Dimensional Properties of Axial Induction Motors. IEEE Trans. Energy Convers.
1986, EC-1, 137–144. [CrossRef]
38. Pyrhönen, J.; Jokinen, T.; Hrabovcová, V. Design of Rotating Electrical Machines; John Wiley & Sons: West Sussex,
UK, 2008; ISBN 9780470695166.
39. IEEE. IEEE Standard Test Procedure for Polyphase Induction Motors and Generators; IEEE Std 112-2017 (Revision
of IEEE Std 112-2004); IEEE: Piscataway, NJ, USA, 2018; pp. 1–115. [CrossRef]
40. Yamazaki, K.; Haruishi, Y. Stray load loss analysis of induction motor comparison of measurement due to IEEE
standard 112 and direct calculation by finite element method. In Proceedings of the IEEE International Electric
Machines and Drives Conference, IEMDC’03, Madison, WI, USA, 1–4 June 2003; Volume 1, pp. 285–290.
41. Levi, E.; Lamine, A.; Cavagnino, A. Impact of stray load losses on vector control accuracy in current-fed
induction motor drives. IEEE Trans. Energy Convers. 2006, 21, 442–450. [CrossRef]
42. Somaloy® Quick Guide. Available online: https://www.hoganas.com/globalassets/download-media/
sharepoint/brochures-and-datasheets---all-documents/somaloy-quick-guide_july_2015_1122hog.pdf
(accessed on 20 September 2020).

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

51
energies
Article
Design and Performance Assessment of a Small-Scale
Ferrite-PM Flux Reversal Wind Generator
Bharathi Manne 1, *, Malligunta Kiran Kumar 1 and Udochukwu B. Akuru 2,3, *
1 Electrical and Electronic Engineering, Koneru Lakshmaiah Educational Foundation, Guntur 522502, India;
[email protected]
2 Department of Electrical Engineering, Tshwane University of Technology, Pretoria 0183, South Africa
3 Department of Electrical Engineering, University of Nigeria, Nsukka, Enugu State 410001, Nigeria
* Correspondence: [email protected] (B.M.); [email protected] (U.B.A.);
Tel.: +91-70-755-50797 (B.M.)

Received: 4 September 2020; Accepted: 29 September 2020; Published: 23 October 2020

Abstract: Currently, there is increasing research interest in harnessing wind energy for power
generation by means of non-conventional electrical machines e.g., flux-reversal machines. The flux
reversal machine is usually designed using scarce rare–earth permanent magnet material which may
be unattractive in terms of machine cost. In this study, an attempt is made to re-design the flux reversal
machine with non-rare-earth ferrite permanent magnet for wind energy applications. Because these
machines possess high cogging torque, which results in vibration and noise, that are detrimental to the
machine performance, especially at low speeds, a novel combined skewed and circumferential rotor
pole pairing method is developed. The proposed cogging torque reduction method is implemented
in 2-dimensional finite element analysis modeling and comparatively analyzed with other existing
stand-alone methods viz., skewing, and rotor pole pairing. The results show that the proposed
method led to 94.8% and 71% reduction in the cogging torque and torque ripple compared to the
reference generator, respectively. However, the calculated torque density is reduced by 13%. Overall,
the electromagnetic performance of the proposed ferrite PM machine exhibits desirable qualities as
an alternative design for the direct drive wind generator.

Keywords: cogging torque; ferrite PM; finite element analysis; flux reversal machine; non-rare earth;
wind energy

1. Introduction
The design of permanent magnet (PM) machines keeps growing significantly by the day due
to advantages of high-power density, high-efficiency, and torque density, useful for variable speed
drives in wind energy applications. The need for increased use of wind energy generation is well
established, due to its availability throughout the day and the possibility for large megawatt generation.
Globally, installed wind power capacity has grown exponentially in recent years as shown in Figure 1,
captured from the recently released Renewables 2020 Global Status Report renewables figure [1].
Wind power generation cost and reliability are critical factors that have to be considered in wind
generator designs, and as such attention is usually focused in the direction of the existing drive
concepts [1]. The various wind generator drive concepts are high-speed (HS), medium-speed (MS),
and low-speed (LS) as characterized in Table 1 [2].

Energies 2020, 13, 5565; doi:10.3390/en13215565 www.mdpi.com/journal/energies

53
Energies 2020, 13, 5565

Figure 1. Global total installed capacity [1].

Table 1. Characteristics of wind turbine drive-trains [2].

Parameter HS MS LS
Speed margin 600–2000 r/min 40–600 r/min 4–35 r/min *
Mass Lightest Intermediate Heaviest
Size Smallest Intermediate Largest
Gearbox presence Yes (3G ¤ ) Yes (1G/2G) Absent
Generator type IG/SG § IG/SG SG
Mechanical losses High Intermediate Lowest
Electrical losses Lowest Intermediate Highest
Cost Gearbox # Intermediate Generator #
* Depends on operating power level. ¤ G represents the stage of the gearbox. § IG = induction generator,
SG = synchronous generator. # highest cost.

PMs are classified into rare-earth (RE) and non-rare-earth (NRE) materials. In recent years,
the unaffordability of RE PM prices forced the electrical machine designers to focus on inexpensive
NRE materials like ferrite PMs and these materials are more desirable for the industry [3]. The trending
flux-reversal machine (FRM), which is a pre-dominantly PM type machine, has been reported in some
studies for rare-earth PM excited wind energy applications [4,5] and direct-drive drive-trains [6,7].
A direct-driven wind turbine generator is important since it generates electricity through the wind
turbine drivetrain by eliminating the gearbox. The advantages of eliminating mechanical gearboxes
from the wind generator system include a reduction in the installation costs due to a lesser number of
components, lower energy costs due to a reduction in losses, and reduction in maintenance costs due
to the simplistic design [6].
Depending on the variation in the rated speed of the wind turbine, direct drive is in the range
of 120–500 r/min, with a corresponding generator power range of 0.5–3 KW [6]. The conventional
synchronous and induction generators are not suitable for LS and MS applications [7]. The majority of
LS and MS applications are handled by mechanical drives (mechanical gear and high-speed motor).
For LS, existing generators require a large number of stator slots and poles, resulting in a machine with
a large air-gap diameter [7].
Meanwhile, ferrite PM materials offer less expensive PM options for machine design but are more
demagnetization prone due to the low coercivity compared to RE PM materials [8]. The affordability
of NRE materials is mainly due to its availability as shown in Figure 2 [9].

54
Energies 2020, 13, 5565

Figure 2. Global permanent magnet (PM) materials market capacity, 2013–2024 [9].

The FRM machine belongs to the category of double salient permanent magnet (DSPM) machines
with PMs on the stator pole surface. Most of the PM machines employ interior rotor PM or surface
mounted rotor PMs. Here, in this PMs are in the rotating part, which may cause demagnetization
problems, limitations on mechanical instability, and poor thermal dissipation. These structures are not
favorable for HS and MS applications [10,11]. To this end, stator-mounted permanent magnet (SMPM)
machines were introduced, they are basically, flux switching machines (FSM) and flux reversal machines
(FRM). These machines generally have similar features such as easy rotor structure, short winding
terminals, good at thermal conditions [12]. FSM as a single-phase flux switch alternator and was
introduced by Rauch and Johnson in 1955. It exhibited poor rotor volume usage causing stator
vibrations and difficulty in stator fabrication [13]. To overcome these difficulties in the manufacturing
process and to improve the torque (power) density, Deodhar et al. introduced single-phase FRM in
1997 for automobile applications to replace standard claw pole alternator [14,15].
The basic 3-phase FRM of 6/8 slot-pole combination with concentrated winding was introduced
by Wang et al. [16,17]. In this machine, the design was optimized from single phase to 3-phase
to ensure maximum PM flux-linkages in the stator winding, lighter PM, and low cogging torque
(CT). However, this type of PM machines exhibit high torque ripple and lead to vibration and noise.
The main cause of the torque ripple is the cogging torque (CT). The CT is mainly position-dependent
and load-independent torque, caused due to the PMs, it mainly deteriorates the machine performance
especially at LS and MS. More and more researchers dedicated themselves to suppress the CT in
RE FRM while maintaining all other machine performances [18–22]. As seen from the literature,
the CT reduction in the FRM plays major role in-terms of ripple-torque. Most of the CT minimization
methods for FRM, published in the literature are some auxiliary techniques like notching, pole pairing,
and skewing [18–21]. These techniques may be limited for the FRM as nothing serious has been
reported for NRE variants. Besides, FRM requires less number of stator slots and a large number of
rotor poles, which qualifies them for direct-driven wind energy generators [5,7].
The main contribution of this paper is to reduce the cogging torque and torque ripple by
a combination of techniques in the proposed ferrite PM FRM while maintaining respectable
electromagnetic performance compared to the basic machine as necessary for direct-drive wind
generator applications. Thus, an overview of the FRM technology including working principle,
machine capabilities using a flux-mmf diagram (FMDT), as well as design features based on different
RE topologies, are discussed in Section 2. In Section 3, the finite element analyses (FEA) on two
existing stand-alone cogging torque minimization methods, as well as the novel method combining
these two methods are undertaken. The power generating performance of all the four generators
considered in this paper in terms of output voltage, power density, voltage regulation, efficiency at the

55
Energies 2020, 13, 5565

rated condition, and overload/speed capabilities are discussed and compared by FEA. In Section 4,
some concluding remarks are given.

2. Overview of RE-PM FRMs

2.1. Structure and Configuration of RE-PM FRM


An early example of 6/4 pole double salient non-rotating PM type motor whose PMs are installed in
the stator yoke is illustrated in Figure 3. This double salient structure enables the superior performance
of torque production, small frame sizes and qualifies the motor as a potential alternative to existing servo
drives, variable speed drives, as well as for satisfying the increasing demand in future automobiles.
The experimental test results have been encouraging, demonstrating twice the output capability,
with higher efficiency and power density when compared with the induction motor [11]. A new
brushless double-salient 2/3 pole FRM, as depicted in Figure 4, is designed, analyzed, and fabricated
based on a single-phase prototype machine for high-speed generator. FMDT is employed to analyze
the qualitative performance of FRM with other types of brushless machines [15].

Figure 3. 3-phase, six/four-pole double salient permanent magnet (DSPM) machine [11].

Figure 4. 1-phase, two/three-pole stator-PM generator [15].

Three-phase 6/8 FRM has been introduced by Wang as depicted in Figure 5, where the magnetic
field distribution, self and mutual inductances, cogging torque variations with rotor positions are
analyzed through 2D FEA [16,17]. In addition, 2-phase, 3-phase, and 5-phase pole combinations with
suitable rotor skewing techniques are documented to minimize the cogging torque. The machine

56
Energies 2020, 13, 5565

capabilities in terms of low rotor inertia, low electrical time constant, and high torque density, as proven
by FEA, had been discussed.

Figure 5. 3-phase, eight/six-pole flux-reversal machine (FRM) [16,17].

Another example of FRM for the servo drive application is depicted in Figure 6 with 12/28 poles [23].
The design specifications and operation of FRM for low-speed, high-torque applications are explained
in [23]. The inset type PM structure is proposed to reduce the flux fringing of 12/40 pole FRM for
servo drives (LS). Through FEA, it is shown that this configuration has achieved high efficiency,
high torque density, and less than 3% torque ripple with three-phase sinusoidal vector control. Already,
the candidature of the FRM for wind applications is growing as shown by some studies [24,25].
A 3-phase, 6/14 pole, 2.4 kW, 214 r/min outer rotor FRM has been introduced by D. S. More [24].
Through experimentation, it is concluded that inner rotor FRM has 1.25 times lower power density
than outer rotor FRM. These two types of machines are depicted in Figures 7 and 8. Power density
comparisons have been made in [25]. Compared with other types of DSPM, it is found that FRM has
higher power density [5] for fractional-slot concentrated winding (FSCW) of FRM and permanent
magnet synchronous machine (PMSM). Through experimentation, it is found that FRM has 1.5 times
higher power density. Both the machine efficiencies are approximately the same [25]. To further
improve the power density of FRM, a full-pitched winding (FPW) is incorporated in the stator of the
FRM as shown in Figure 9 [26].

(a) (b)

Figure 6. Various pole and PM arrangements of the FRM machines [23]; (a) PMs on stator pole;
(b) inset PMs on stator.

57
Energies 2020, 13, 5565

Figure 7. A typical 3D Structural view of 6/14 inner pole rotor of FRM [24].

Figure 8. Cross-section of 3phase, 6/14 outer pole rotor FRM [24].

Figure 9. Structural-view of 6/14 FRM [25,26]; FPW: Full-pitched winding, CW: Concentrated winding.

Comparing different topologies of the FRM, that is, the consequent-pole permanent magnet
(CPM) topology with the SMPM topology, as illustrated in Figure 10, it was concluded that CPM
topology improves the torque performance and reduces the magnetic volume [27]. To reduce the cost
of PMs, consequent-pole transverse-flux permanent magnet linear machine (TFPMLM) is proposed,
partially replacing PM Poles by soft magnetic iron in [27,28].

58
Energies 2020, 13, 5565

(a) (b)

Figure 10. Typical topologies of FRM: (a) consequent-pole permanent magnet (CPM)-FRM);
(b) stator-mounted permanent magnet (SMPM)-FRM [27,28].

Introducing a small space-gap between the adjacent PMs belonging to the same stator pole
shoe, as illustrated in Figure 11, has helped to improve the phase back-emf and cogging torque [29].
FRM designed with soft magnetic composite materials for fans exhibited a significant increase in
efficiency while producing high power density and reducing the usage of PMs compared to the
existing PM type machines [30]. Based on electromagnetic compatibility, high-speed FRM topology
has been proposed for the angular grinder as shown in Figure 12 and experimental results show that
the efficiency of the developed machine is higher than that produced in induction and brushless type
motors [31].

Figure 11. 3-phase, 6-stator-pole/8-rotor-pole FRM [29].

Figure 12. Cross-section of 1-phase, FRM motor [31].

59
Energies 2020, 13, 5565

Comparisons with the different configurations of the FRM showed that the stator flux linkages of
the 6/14 FRM is doubled compared to that of the 12/16 pole FRM as illustrated in Figure 13 [32]. This is
because in the 12/16 pole FRM, the stator teeth surface occupies around 2/3 of the inner stator surface
and 1/3 of the stator inner surface is wasted. Accordingly, it causes high cogging torque (CT), vibrations,
and acoustic noise. To overcome all the constructional issues of the 12/16 pole FRM, Dmitrievskii
introduced the inner stator surface 12/10 pole FRM in 2018 as shown in Figure 14 [32]. Acoustic noise
and CT are further reduced by this configuration for wind applications.

Figure 13. Schematic representation of 12-stator-slot/16-pole-rotor of FRM [32].

Figure 14. Configuration of a 12-stator-slot/10-pole-rotor of FRM [32].

A 3 kW, 200 r/min, 48/46 pole FRM has been projected for direct-drive wind energy [32] and
is depicted in Figure 15, FEA calculations are done for FRM and PMSM with the same machine
dimensions, power ratings, and speed. It is concluded that core volume is reduced by 25.6% and
PMs required for FRM is five times less than that for PMSM and correspondingly, the cost of the
active material like the PMs becomes twice as low as for PMSM. Then, the number of poles of FRM is
higher, the frequency of FRM is thrice as high at the same speed of both the machines, and hence FRM
is best suitable for variable speed applications than PMSM. The FRM cooling is simpler, and it can
run at higher ambient temperatures. The FRM efficiency is 2.3% higher than the PMSM for gearless
wind energy applications [31]. More [5,6] and Pellegrino [7] introduced fictitious electrical gear of
6/14 FRM for direct drivetrains i.e., low-speed high torque applications. It is concluded with fabrication
results that FRM can be considered as PMSM with an inbuilt gearing effect [5,6]. The various design
parameters influencing FRM performance are also analyzed [7]. A new proposed FRM structure
is examined to improve the CT profile by changing the PM thickness and rotor side geometrical
parameters for wind generators [33].

60
Energies 2020, 13, 5565

Figure 15. 3-phase, 48/46-pole FRM generator for the direct drive [32].

2.2. Basic Principle, Design Topologies, and Performance of FRM


FRMs are hybrid machines combining the advantages of a switched reluctance machine (SRM)
and DSPM into one machine. Both the rotor and stator poles have double salient structures. For this
non-conventional type of machine, the operational capabilities are analyzed by the flux-mmf diagram
technique (FMDT) and implemented using FEA by Ion Boldea in 1996 [14,15]. FMDT has its origin
in the ψ–I diagram and can predict the periodic variation of phase mmf and flux variations over an
electrical cycle. The FMDT for three double salient type machines for comparisons is illustrated in
Figure 16. The area enclosed by the flux-mmf loop is the average energy converted over an electrical
period and it specifies the average electro-magnetic torque produced over a rotor movement. From this
comparison, it is seen that the SRM has uni-polar mmf; phase-flux variations and energy conversion
loop are limited to the first quadrant only whereas the DSPM has bipolar mmf and uni-polar phase-flux,
with the energy conversion loop limited to two quadrants. However, the FRM has bipolar mmf and
phase-flux variation with respect to rotor movement, the energy conversion loop covers the four
quadrants; meaning that control is possible for all four quadrants. The typical 2D and 3D cross-section
of FRM of 6/8 poles are illustrated in Figure 17. The flux and flux density distribution in the machine
at aligned and un-aligned positions of the rotor are illustrated in Figure 18. The variation of mmf
and phase-flux with respect to rotor displacement are shown in Figure 19. The principle of FRM is
variable flux linkages, inducing an emf that interacts with alternating armature current as seen in
Figure 18. The ideal variation of mmf and phase flux with respect to the rotor displacement is seen in
Figure 19. The field excitation provided by the PMs, armature winding flux linkages are modulated by
the variation of the magnetic circuit reluctance, as rotor displaces, in such a way that bipolar induced
emf are produced without rotating the PMs [15].

Figure 16. Comparisons with three double salient type machines [14].

61
Energies 2020, 13, 5565

Figure 17. A typical 3D and 2D structures of designed FRM.

(a) (b)

(c) (d)

Figure 18. Flux-distribution of six/eight-pole FRM: (a) rotor position at 0o ; (b) rotor position at 11.25o
ACWD; (c) rotor position 22.5o ACWD; (d) rotor position at 11.25o CWD.

Figure 19. Variation of mmf and phase-flux variations with respect to rotor displacement [15].

In the generator case, when the rotor is driven by the prime mover as in Figure 18a, the equilibrium
position of the rotor poles show they are unaligned with stator poles. At that position, there are no
flux linkages with the coils. Only the flux setup by the magnets circulates completely within each
stator teeth (point “a” in Figure 19). In Figure 18b, the rotor is driven to 11.25o in an anti-clockwise
direction (ACWD) then the rotor poles overlap with the stator pole magnets and flux creates the
path from upper PMs, lower PMs, and stator back core iron. Phase flux is extreme in this position at

62
Energies 2020, 13, 5565

(point “b” in Figure 19). In Figure 18c, again at equilibrium position which is displaced from the first
equilibrium position by 22.5o no flux links with the coils (point “c” in Figure 19). Further movement of
the rotor leads to the position in Figure 18d of 11.25o clockwise direction (CWD), where phase flux is
again extreme in the opposite direction to that of the first alignment position (point “d” in Figure 19).
Linear bipolar phase flux variation induces a total induced emf, eO . According to the Faraday’s law
of electromagnetic induction (1), a change in flux linkage produced by ψm (field source), at a given
electrical speed ωe in rad/s, the emf induced eO is given as

dψm
eO = −ωe (1)

Table 2 shows the evaluated dimensions of a FRM through the sizing equation technique [23].
Based on these dimensions, FRM has been analyzed in 2D FEA and by using the magneto-static solver
analyzed the flux density distribution and cogging torque. The meshed plot and no-load flux-density
distribution of the modeled machine are shown in Figure 20. The magneto-static cogging torque is
shown in Figure 21. It is important to note that FRM exhibits fault tolerance ability because the mutual
inductance magnitude value is less than one-fiftieth of the self-inductance of the phases which indicates
that natural isolation between the phases. The variation of the inductances with respect to the rotor
displacement is small. Here the mutual inductance 4 mH and self-inductance is 0.61 H. Therefore,
the reluctance torque is insignificant.

Table 2. Dimensional details of non-rare-earth (NRE) FRM.

S. No Parameters Value Symbol


magnet thickness 3 mm hpm
stator-pole span angle 42.6o βs
stator-pole height 15 mm hps
outer diameter of stator 129 Ds
Stator stator arc span 27.8 mm τps
stack length 400 mm lstk
no. of turns/coil 176 Nph
air gap 0.5 mm g
magnet remanence 0.4T (Sr-Fe) 0.4T (Sr-Fe)
outer diameter of rotor 72 mm Dr
Rotor rotor arc span 12.3 mm τpr
rotor pole span angle 21o βr
rotor pole height 17 mm hpr

(a) (b)

Figure 20. FEA evaluation of no-load behavior of FRM: (a) mesh plot; (b) flux density map.

63
Energies 2020, 13, 5565

Figure 21. Cogging torque cycle under magneto-static no-load (MSNL).

2.3. Design Topologies


A general rule regarding the number of rotor poles Nr and the number of stator poles Ns and
choosing “p” for the number of phases of FRM is [17]:

Ns p
= (2)
Nr p+1

Hence Ns /Nr = 3/4, 6/8, 12/16, etc., are prevalent for 3-phase FRM machines (2).
If a number of pairs of PMs are given, then [23]:
 
np + 1
Ns = Nr (3)
3

where np = number of PMs pair poles. Various possible configurations of the FRM are shown in Table 3.
For 6/8 FRM configuration accompanied one pair of PMs in each stator pole (3). The no-load flux
density plot of 6/8 pole FRM is illustrated in Figure 20. In addition, the number of possible PMs on the
stator surface and the number of effective poles corresponding to the flux patterns are given in Table 3.
Table 3 shows that as the number of rotor poles increases, speed decreases. Increasing the rotor pole
number means increasing the energy cycles per revolution. Thus, the reluctance torque is negligibly
small, even though the CT is still slightly high for LS applications. With the proper design and control,
the cogging torque component can be made relatively small. The number of poles corresponding to the
effective flux patterns, speed, and gear ratio for different FRM topologies for LS applications are given
in Table 3. A summary of the various FRM topologies highlighted so far is evaluated from ten points
of view with reference to wind turbines in Table 4. The performance marks are given in a percentage
scale. In Table 4, these percentage values are represented in form of their fractional values.
Table 3. Various possible configurations of PM FRM.

Machine Number of Effective Number of Speed for 50 Cogging


Type Flux Patterns Magnets Hz (r/min) Torque Cycle
6/8 2 12 375 15o
6/14 2 24 214 8.57o
12/16 4 24 188 7.5o
12/10 4 24 300 6.0o
12/28 4 48 108 4.2o
12/40 4 60 75 3.0o
48/46 4 96 65.3 0.33o

64
Table 4. Comparisons of various topology rare-earth (RE)-PM FRMs for wind energy applications.

Figure 4 Figure 5 Figure 6a Figure 6b Figure 7 Figure 8 Figure 9 Figure 10a Figure 10b Figure 11 Figure 12 Figure 13 Figure 14 Figure 15
Properties
[15] [16,17] [23] [23] [24] [24] [25,26] [27,28] [27,28] [29] [31] [32] [32] [32]
Constructional
1 1 1 0.8 0.8 0.6 0.4 0.4 0.4 0.8 0.1 0.8 0.8 0.6
simplicity
Power Density 0.6 0.8 0.8 0.8 0.8 1 1 0.8 1 0.8 0.6 0.6 0.8 0.8
Energies 2020, 13, 5565

Usage of Rotor Volume 0.8 0.8 0.8 0.6 0.8 0.8 0.6 0.8 1 0.8 0.8 0.4 0.8 1
Low material&
1 0.8 0.6 0.4 0.8 0.6 0.6 0.4 0.4 0.8 1 0.6 0.6 0.6
manufacturing cost
Low CT 0.8 0.8 0.6 0.6 0.8 0.8 0.8 0.6 0.6 0.8 0.8 0.4 0.4 0.6
High Temperature
0.8 0.8 0.8 0.6 0.8 0.8 0.8 0.8 1 0.8 0.8 0.6 0.6 0.8
operation
Fault tolerance 1 0.8 0.8 0.6 0.8 0.8 0.8 1 0.8 1 0.6 0.6 0.8 0.8
Easy replacement of
0.8 0.8 0.6 0.6 0.8 0.6 0.6 0.8 0.8 0.8 0.6 0.6 0.8 0.6
faulted coils
Mechanical rigidity 0.8 0.8 0.6 0.6 0.8 0.6 0.8 0.6 0.6 0.8 0.6 0.6 0.6 0.8
Flux Leakage 0.8 0.8 0.6 1 0.8 0.8 0.8 0.8 0.8 0.8 0.6 0.6 0.6 1

65
Energies 2020, 13, 5565

3. Analytical Calculation of Cogging Torque and Torque Ripple


The 2-dimensional governing expression of PM FRM can be expressed in magnetic vector potential
A as [34]:      
∂ 1 ∂Az ∂ 1 ∂Az ∂My ∂Mx
+ = −Jo − − (4)
∂x μ ∂x ∂y μ ∂y ∂x ∂y
where Az = vector potential in z, Jo = current density, M = magnetization of PM.
The flux linkage λ are calculated per phase from the average vector potential ‘A’ (4) over each
winding cross-section as ⎛ ⎞
⎜⎜ ds  ds2 ⎟⎟⎟
λ = ⎜⎜⎜⎝ A1 1 − A2 ⎟⎟l (5)
s1s1
s2 ⎠ stk
s1

where lstk is stack length; s1 and s2 are the total area of Nph -turns per phase with the winding carrying
the negative and positive currents, respectively. The flux linkages of each phase of 3-phase FRM
machines can be evaluated (5).
FRM mainly suffers from unfavorably large CT due to its double-salient construction and PMs,
resulting in undesirable vibration and noise especially for LS and MS wind energy applications. At high
speed, this effect is negligible [18]. The CT does not add to the average electro-magnetic torque,
but only affects the pulsations in the torque. It is also known as self-aligning torque or no-current
torque. It occurs mainly in PM machines. It is the torque due to the interaction between the PM and
the stator slots. When there is no current in stator windings, as the saliency nature of rotor and stator
with uneven permeance in the air gap produces the cogging torque [18]. The no-load open-circuit
flux density distribution is obtained from Equation (6), based on the mmf-permeance model in the
airgap [19,20].
B(α, θ) = Fpm (α) ∧ (α, θ) (6)

where ∧(α, θ) and Fpm (α) are permeance in the air gap and the mmf excited by PMs, respectively. α is
the rotor position along the air gap circumference, θ is rotor position, w is the total energy stored in the
air gap and υ is the volume of the air-gap.
The analytical expression of CT of FRM machines can be obtained [20] from the instantaneous
torque for every cycle of rotor displacement by means of the rate of change of co-energy with respect
to the rotor movement as shown in Equation (7).
The 2D FEA is used to predict the overall CT as
 " 
Tcog = − ∂w
∂θ

= − ∂θ 1
2μo B2 (α, θ) dυ
 "  (7)

= ∂θ 1
2μ F2pm (θ) ∧2 (α, θ)dυ
o

The cogging torque is calculated when the machine is at a standstill position and when there is no
current in the coils of the stator (i = 0). It is supposed that each section plane along axial direction
has the same magnetic field distribution (6). The cogging torque of FRM is obtained by having the
calculated value multiplied by the total machine length or stack-length [18]. The overall cogging torque
mainly depends on airgap permeance and PMs mmf functions (7).
The analytical expression of total cogging torque for conventional stator active PM brushless FRM
can be expressed as [20]:
 
R22 − R21 La Nr π 

Tcog = mFpm(mNr /Ns ) ∧m sin(mNr θ) (8)
4μo
m=1

66
Energies 2020, 13, 5565

where La = effective stack length, R1 and R2 are the outer rotor radius and inner stator radius, ∧m is
Fourier coefficient, and m is satisfied as the following expression is given as

kNs
m= k = 1, 2, 3 . . . .. (9)
GCD(Ns , Nr )

where GCD(Ns , Nr ) is the highest common factor of Ns and Nr . The Equations (8) and (9) can be
appropriate for FRMs of any slot-pole configurations [20].
The torque ripple factor krp is evaluated as
# $
Tmax − Tmin
krp = × 100 % (10)
Te

3.1. Cogging Torque Reduction Techniques for PM FRM


Many researchers studied and developed different design techniques to minimize the CT of
PM machine structures [19]. The most accepted techniques to minimize CT is to optimally vary the
dimensions of the machine design such as rotor skewing, rotor teeth pairing, chamfering the rotor
and stator PM thickness, and rotor geometrical modifications. Generally, stator side modifications
consist of tooth pairing, magnet skewing, slot shifting, and dummy slots [35,36]. The stator side
modifications augment the machine size and cost. Thus, the stator modifications are not practical in
the case of PM machines. Then, the rotor side modifications are skewing rotor, chamfering, and radial
pole pairing, and axial pole pairing [18–20]. The required modifications in the rotor than in the
stator are generally used for minimizing cogging torque. Therefore, in this paper, the analysis part is
divided into three subsections, where the first and second are used to report analysis of the existing
CT minimization techniques like skewing and rotor pole pairing (CPOP). In the third subsection,
a novel model constituting, the combining of these two existing techniques (SKCpp) is additionally
investigated. Compared are different methods for electromagnetic performances of a generator with
various rotor side design modifications.

3.1.1. Skew Rotor Design


Skewing is the most renowned and widely used technique to minimize the CT effects in PM
machines. Skewing can be done in either stator or rotor. In the FRM, the stator has magnets while the
rotor is composed entirely of laminations. So, rotor skewing is more feasible. Furthermore, when the
rotor is skewed with the optimal skewing angle, the CT is effectively minimized. The skewed rotor
yields even permeance in between the stator and rotor irrespective of the rotor displacement [17].
In this study, slicing or stepped skew technique has been incorporated into the rotor structure and
the rotor is sliced into five segments along the length. For various skew angles (5o , 10o , 15o , and 20o ),
the CT effect is analyzed by 2D FEA as shown in Figure 22. The CT variation with rotor position has a
cycle of 2π/Ns . The CT cycle is 15o , which agrees with FEA simulation as in Figure 22. Based on the
simulation 15o skew angle effectively minimized the CT compared with the other angles. FEA shows
that the CT is decreased by 92.2%. Nevertheless, this method leads to some reduction of flux linkages,
induced emf, and the corresponding reduction of power density.

67
Energies 2020, 13, 5565

Figure 22. The effect of cogging torque for various skew angles.

3.1.2. Circumferential Rotor Teeth Pairing (βr ) Method or Rotor Pole Pairing (CPOP)
The rotor teeth pairing method is also considered to minimize the CT of FRM. The CT waveform
varies with the rotor tooth width βr . This method of circumferential or rotor teeth pairing method
employs two different pole widths in the rotor design that can be applied. The variable magnetic
reluctance of the rotor and the air gap minimizes the amplitude of the CT. Based on the Fourier
series expansion [18,20], by adjusting the rotor pole width with respect to the PM width with these
combinations the optimal rotor tooth width has to be selected as 19o and 21o . The rotor poles are
designed into two different pole arc widths, these two arcs are oppositely employed as illustrated in
Figure 23. The overall cogging torque got reduced as verified by FEA. This technique reduced the CT
by 14% of the original value in transient no-load. The overall CT can also be minimized.

Figure 23. Cross-section of teeth paired rotor.

3.1.3. Combined Auxiliary Model of FRM


The skewing and rotor pole pairing methods are explained above; this paper attempts to combine
these two methods to obtain the performance of FRM better than any one of the two. This combined
method is the possible combinations of more than two existing CT reduction methods. There are
various combined methods are studied for PM machines. For example, the combination of rotor
magnet skewing with teeth notching by B. Zhang [21] and the combined pole and slot number by F.
Yusivar [37]. The 3D structural view of SKCpp is depicted in Figure 24. This method of combining the
existing two methods is a novel method and is being introduced for the first time. By this method,
the CT is reduced by 94.8% of the original value in Transient (TR) no-load by 2D FEA. Compared to

68
Energies 2020, 13, 5565

the conventional skewing, rotor teeth-pairing, the SKCpp also effectively minimized the cogging
torque while maintaining all other electromagnetic performances as in the basic model.

Figure 24. 3D Structural view of SKCpp.

3.2. Basic Performance Evaluation


Some evaluations are carried out to examine the generator performances for each rotor design in
terms of flux linkages, induced voltages, and cogging torque.

3.2.1. Flux Linkage


Under the open-circuit condition, the flux from PMs is investigated with Equation (5).
A 180o rotation of the rotor, with step time as 0.1 ms and simulated time of 80 ms results in four
cycles of 50 Hz frequency and 800 simulations. Figure 25 clearly shows that CPOP has the highest
flux-linkages amplitude that emerges from the PM compared to other generators, while skew has
the lowest flux linkage compared to the other generators. For basic, skew, CPOP, and combined
auxiliary models (SKCpp), the measured maximum flux linkages are 0.41 Wb, 0.327 Wb, 0.433 Wb,
and 0.348 Wb, respectively.

Figure 25. Flux linkage waveforms comparison.

3.2.2. Induced Voltage


Further, under open circuit condition (TR-no-load), the rotor is rotating at a prescribed speed
of 375 r/min. The induced emfs of the four-rotor designs are analyzed and compared in Figure 26.

69
Energies 2020, 13, 5565

As expected, the highest magnitude is measured with CPOP as 118.4 V followed by basic, SKCpp and
skewing with 112 V, 109.5 V, and 101.7 V, respectively.

Figure 26. Induced voltage waveforms comparison at transient no-load.

3.2.3. Cogging Torque (CT) Reduction


The CT as applied to PM machines is the torque effort due to magnets only viz., no load current
present. The CT leads to poor position control, vibration, noise, and reduction in generator performance.
By setting rotor speed at 375 r/min, and one complete electrical cycle of 180o , CT for various rotor
configurations is plotted as shown in Figure 27. This shows the conventional (basic) rotor model has
the highest peak-to-peak CT value of 1.82 Nm followed by the values for skewed, rotor teeth pairing
(CPOP) and SKCpp models as 0.14 Nm, 1.56 Nm, and 0.09 Nm, respectively. As compared to the
basic model, skew and SKCpp gives that highest CT reduction. In Table 5, the reduction in CT at
TR-no-load is compared quantitatively, which indicates that the generator acoustic noise is lowest with
proposed SKCpp.

Figure 27. Comparisons of cogging torque effect for various methods at transient no-load.

70
Energies 2020, 13, 5565

Table 5. Comparison table of reduction of CT for NRE PM FRM at TR-no-load.

Cogging Torque Reduction


Rotor Model Cogging Torque (Nm)
(%)
Conventional (6/8) 1.283 0
Skew 0.1438 92.2
CPOP 1.5625 14.5
SKCpp 0.0940 94.8

3.2.4. Torque Ripple Reduction Analysis (TR-Load)


The characteristics of the torque ripple lead to pulsating torque which results in noise and vibration,
impacting negatively on the machine performance under load. The torque ripple factor is evaluated
by Equation (10) and the torque ripple for different rotor modification techniques at the rated load is
evaluated as shown in Figure 28. It shows that the basic model has the highest torque ripple factor
value of 18.6 % followed by the values of skewing, CPOP, and SKCpp at 6.73%, 17.4%, and 5.5%,
respectively. In comparison to the basic model, the CPOP, and skew, the SKCpp model shows the
highest torque-ripple reduction. In Table 6, the reduction in torque ripple at TR-load is compared
quantitatively with the conventional design as the reference.

Figure 28. Comparisons of a torque ripple effect for various methods at transient load.

Table 6. Comparison table of reduction of torque ripple for NRE-PM FRM at TR-rated-load.

Rotor Model Average Torque, Te (Nm) Torque Ripple Reduction (%)


Conventional (6/8) 19.8537 0
Skew 16.7142 64
CPOP 21.1553 8
SKCpp 18.419 71

3.3. Power Generating Performance Comparison


The generating operating point performances including output voltage against load current,
voltage regulation, power curve in terms of varying load current and generator speed, as well as loss and
efficiency curves of the four generators working are analyzed by FEA simulation. The four generators
are designed to work under symmetrical resistive loads. Hence, based on the rated voltage of each
phase and rated power, the calculated rated load resistance, Rn , is 25.3 Ω.

71
Energies 2020, 13, 5565

Figure 29 illustrates the variation curves of the output voltage versus phase current—-so-called
overload capability curves—-while the voltage regulation is presented in Figure 30. One can see
from Figure 29 that the CPOP is presented with as the machine with the highest overload capability,
while basic, SKCpp and skew follow in that order. In Figure 30, as load current increases, the skew is
presented with the highest voltage regulation, followed by SKCpp, basic, and CPOP. Lower percentage
values in Figure 30 is an indication that there is a less voltage variation when the load changes in
the generators.

Figure 29. Output voltage against phase current @ 375 r/min.

Figure 30. Voltage-regulation against phase current (rms) for each rotor designs @ 375 r/min.

The generator load profiles are exhibited by varying the phase current against the output power
for each rotor design as shown in Figure 31. The overload limits are (4.6875 A, 850.77 W), (4.6875 A,
747.42 W), (4.6875 A, 907.611 W) and (4.6875 A, 825.78 W) for basic, skew, CPOP, and SKCpp,
respectively, which is up to 1.5 times of the rated load. Meanwhile, the efficiencies are compared in
Figure 32, for all generators.

72
Energies 2020, 13, 5565

Figure 31. Output-power against phase current (rms) for each rotor designs @ 375 r/min.

Figure 32. Efficiency against phase current (rms) for each rotor designs @ 375 r/min.

In terms of speed capability, the open circuit EMF increases with speed, which translates to
increasing power density. The speed capability of the generators are analyzed when connected with
a symmetrical rated load while varying the mechanical speed as shown in Figure 33. The output
power of the skewed rotor generator becomes easily saturated and decreases with the increase in
the speed of the rotor. The core loss and efficiency curves against rotor speed are also plotted as
shown in Figures 34 and 35, respectively. As expected, an increase in speed, increases the core-loss
magnitudes. As the rotor speed increases, the core losses become distinct and it is observed that SKCpp
presents the lowest value compared with the other generators.
Overall, from the 2D transient FEA simulations, a summary of the performance for all
four generator designs is provided as shown in Table 7. It is interesting to note that the proposed
SKCpp design yields comparatively improved performance characteristics in terms of cogging torque
and torque ripple reduction, as well as efficiency, while CPOP presents better performance with respect
to stability for overload capability and voltage.

73
Energies 2020, 13, 5565

Figure 33. Output-power against mechanical speed (rpm) for each rotor design @ 375 r/min.

Figure 34. Core-loss against rotor mechanical speed (rpm) for each rotor design @ 375 r/min.

Figure 35. Efficiency against rotor mechanical speed (rpm) for each rotor design @ 375 r/min.

74
Energies 2020, 13, 5565

Table 7. Comparison table of all the rotor designs of 6/8 NRE-PM FRM Generators.

Rotor Model Cogging Torque Torque Ripple PowerLimit Voltage Regulation Efficiency
Basic (6/8) Poor Poor Moderate Moderate Good
Skew Moderate Moderate Poor Poor Moderate
CPOP Poor Poor Good Good Good
SKCpp Good Good Moderate Moderate Good

4. Conclusions
In this study, the possibility of using NRE (ferrite) PM excitations to design the flux reversal machine
for direct drive wind turbine generator has been demonstrated for the first time. Various existing RE
PM-FRM structures have been reviewed. In this frame of overview, fourteen RE PM FRM designs
were highlighted and examined in detail for their suitability for wind energy generation. The merits
and demerits of these FRM topologies were quantitatively deliberated in terms of power density,
cogging torque, and torque-ripple.
Thereafter, a 6/8 pole ferrite PM FRM is redesigned with a conventional rotor model, as well as
with skewing and rotor pole pairing, while a new combination of skewing with circumferential
rotor pole pairing (SKCpp) is also designed and proposed for minimizing cogging torque and torque
ripple. The machines are numerically evaluated for wind power generation in 2D FEA. In addition,
the generator performance, in terms of flux linkages, induced voltages, over-load and over-speed
capabilities, as well as efficiency, are investigated and discussed. The results obtained through 2D FEA
simulations show that the proposed SKCpp technique as the best design approach for reducing the
cogging torque and torque ripple effect in ferrite PM FRM, with a percentage reduction of 94.8% at
no-load and 71% under load, respectively. However, the average torque is reduced by 13%. Overall,
the potential of the proposed ferrite PM FRM for wind power generation is clearly demonstrated for the
reduction of cogging torque and torque ripple and also for efficiency, overload, and speed regulation.

Author Contributions: Conceptualization, B.M. and U.B.A.; methodology, B.M. and U.B.A.; software, B.M.;
validation, B.M. and U.B.A.; formal analysis, B.M. and U.B.A.; investigation, B.M.; resources, M.K.K.; data curation,
B.M.; writing—original draft preparation, U.B.A. and B.M.; writing—review and editing, B.M., U.B.A. and M.K.K.;
visualization, B.M.; supervision, U.B.A. and M.K.K.; project administration, B.M.; funding acquisition, U.B.A.
All authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding but the APC is funded by Tshwane University of Technology,
South Africa.
Conflicts of Interest: The authors declare no conflict of interest.

Nomenclature
Symbol Stands for
eO Induced EMF (V)
ωe Electrical speed (rad/s)
hpm PM thickness (mm)
τps Stator arc span (mm)
hps Height of stator pole (mm)
g Length of the airgap (mm)
lstk Length of stack laminations (mm)
Ds Outer diameter of stator (mm)
hpr Height of rotor pole (mm)
βr Span angle of rotor pole (o mech.)
βs Span angle of stator pole (o mech.)
Dr Outer diameter of rotor (mm)
La Effective stator stack length
Nph Number of turns per coil
m Number of phases

75
Energies 2020, 13, 5565

Nr Rotor pole number


ψm The field source Flux linkage (Wb)
Ns Stator pole number
np Number of PMs Pairs
Tcog Cogging torque (Nm)
A Magnetic vector potential (T-m)
Jo Current density per phase (A/mm2 )
M Magnetization of PMs (A/mm)
τpr Rotor arc span (mm)
λ Flux linkages per phase (Wb-t)
θ Position of the rotor (o mech.)
w Magnetic co-energy (J)
Fpm MMF due to Permanent magnets (A-t)
B No-load flux density (T)
∧ Permeance (Wb/At)
υ Air-gap volume (mm3 )
μo Permeability of free space (H/m)
s Area of Nph turns
R1 Outer rotor radius (mm)
R2 Inner stator radius (mm)
krp Ripple-torque factor (%)
α Position along the circumference in the airgap
Tmax Maximum Torque (Nm)
Tmin Minimum Torque (Nm)
Te Average Torque (Nm)

Abbreviations
PM Permanent magnet
RE Rare-earth
NRE Non-rare earth
HS High speed
LS Low speed
FRM Flux reversal machine
FSM Flux switching machine
SMPM Stator-mounted permanent magnet machine
CT Cogging torque
FMDT Flux-mmf diagram
FEA Finite element analysis
PMSM Permanent magnet synchronous machine
FSCW Fractional slot concentrated winding
FPW Full pitched winding
CW Concentrated winding
CPM Consequent-pole permanent magnet
TFPMLM Transverse-flux permanent magnet linear machine
SRM Switched reluctance machine
DSPM Double salient permanent magnet machine
TR Transient
CWD Clockwise direction
ACWD Anti-clockwise direction
CPOP Circumferential rotor pole pairing
SKCpp Skewing with circumferential pole pairing

76
Energies 2020, 13, 5565

References
1. REN 21. Renewables 2020 Global Status Report (Paris: REN21 Secretariat). Available online:
http://www.ren21.net/gsr (accessed on 10 October 2020).
2. Akuru, U.B.; Kamper, M.J. Evaluation of flux switching PM machines for medium-speed wind
generator drives. In Proceedings of the IEEE Energy Conversion Congress and Exposition (ECCE),
Montreal, QC, Canada, 2015; pp. 1925–1931.
3. Thomas, J. Getting rare-earth magnets out of EV traction machines: A review of the many approaches being
pursued to minimize or eliminate rare-earth magnets from future EV drive trains. IEEE Electrif. Mag. 2017,
5, 6–18.
4. Arshad, S.; Selvi, V. 2D electromagnetic design of flux reversal generator for low speed wind applications
using finite element analysis. Int. J. Eng. Technol. Sci. Res. 2017, 4, 149–157.
5. More, D.S.; Kalluru, H.; Fernandes, B.G. Comparative analysis of flux reversal machine with fractional slot
concentrated winding PMSM. In Proceedings of the 34th Annual Conference of IEEE Industrial Electronics,
Orlando, FL, USA, 10–13 November 2008; pp. 1131–1136.
6. More, D.S.; Fernandes, B.G. Analysis of flux-reversal machine based on fictitious electrical gear. IEEE Trans.
Energy Convers. 2010, 25, 940–947. [CrossRef]
7. Pellegrino, G.M.; Gerada, C. Modeling of flux reversal machines for direct drive applications.
In Proceedings of the 14th European Conference on Power Electronics and Applications, Birmingham, UK,
30 August–1 September 2011; pp. 10–18.
8. Akuru, U.B.; Kamper, M.J. Investigation of low-cost PM flux switching machine for medium speed geared
wind energy application. In Proceedings of the 25th South African Universities Power Engineering Conference
(SAUPEC), Stellenbosch, South Africa, 2017; pp. 613–618.
9. Akuru, U.B. Design Optimization and Performance Evaluation of Flux Switching Machines for Geared
Medium-Speed Wind Generator Drives. In Proceedings of the IEEE PES/IAS Power Africa Conference,
Cape Town, South Africa, 2018; pp. 925–930.
10. Yuting, G.; Ronghai, Q.; Dawei, L.; Li, J. Torque performance analysis of three-phase flux reversal machines
for electrical vehicle propulsion. In Proceedings of the 2016 IEEE Transportation Electrification Conference
and Expo Asia-Pacific (ITEC Asia-Pacific), Busan, Korea, 1–4 June 2016; pp. 297–301.
11. Liao, Y.; Liang, F.; Lipo, T.A. A novel permanent magnet motor with doubly salient structure. IEEE Trans.
Ind. Appl. 1995, 31, 1069–1078. [CrossRef]
12. Jianzhong, Z.; Ming, C.; Zhe, C.; Wei, H. Comparison on stator-mounted permanent-magnet machines based
on a general power equation. IEEE Trans. Energy Convers. 2009, 24, 826–834. [CrossRef]
13. Rauch, S.E.; Johnson, L.J. Design principles of flux-switch alternators. IEEE Trans. Am. Inst. Electr. Eng. 1955,
74, 1261–1268.
14. Deodhar, R.P. The Flux-MMF Diagram Technique and Its Applications in Analysis and Comparative
Evaluation of Electrical Machines. Ph.D. Thesis, University of Glasgow, Glasgow, UK, October 1996.
15. Deodhar, R.P.; Andersson, S.; Boldea, I.; Miller, T.J.E. The flux reversal machine: A new brushless
double-salient permanent magnet machine. IEEE Trans. Ind. Appl. 1997, 33, 925–934. [CrossRef]
16. Wang, C.; Nasar, S.A.; Boldea, I. Vector control of three-phase flux reversal machine. Electr. Mach. Power Syst.
1999, 28, 153–166.
17. Wang, C.; Nasar, S.A.; Boldea, I. Three phase flux reversal machine (FRM). IEE Proc. Electr. Power Appl. 1999,
146, 139–146. [CrossRef]
18. Zhao, W.Y.; Chen, Y.G.; Shen, Y.B.; Xing, S.Z. Effective methods of reducing cogging torque in flux-reversal
machine. J. Iron Steel Res. Int. 2006, 13, 444–449. [CrossRef]
19. Zhu, X.; Hua, W. An improved configuration for cogging torque reduction in flux-reversal permanent magnet
machines. In Proceedings of the 2016 IEEE Conference on Electromagnetic Field Computation (CEFC),
Miami, FL, USA, 13–16 November 2016; pp. 18–21.
20. Xiaofeng, Z.; Wei, H.; Zhongze, W. Cogging torque suppression in flux-reversal permanent magnet machines.
IET Electr. Power Appl. 2018, 12, 135–143.
21. Zhang, B.; Wang, X.; Zhang, R.; Mou, X. Cogging torque reduction by combining teeth notching and rotor
magnets skewing in PM BLDC with concentrated windings. In Proceedings of the 2008 International
Conference on Electrical Machines and Systems, Wuhan, China, 17–20 October 2008; pp. 3189–3192.

77
Energies 2020, 13, 5565

22. Chu, W.Q.; Zhu, Z.Q. On-load cogging torque calculation in permanent magnet machines. IEEE Trans. Magn.
2013, 49, 2982–2989. [CrossRef]
23. Boldea, I.; Zhang, L.; Nasar, A. Theoretical characterization of flux reversal machines in low-speed servo
drives- the pole-PM configuration. IEEE Trans. Ind. Appl. 2002, 38, 1549–1557. [CrossRef]
24. More, D.S.; Kalluru, H.; Fernandes, B.G. Outer rotor flux reversal permanent magnet machine for rooftop
wind generator. In Proceedings of the Industry Applications Society Annual Meeting 2008 IAS’ 08,
Edmonton, AB, Canada, 25 April 2008; pp. 1–6.
25. More, D.S.; Fernandes, B.G. Power density improvement of three phase flux reversal with distributed
winding. IET Electr. Power Appl. 2010, 4, 109–120. [CrossRef]
26. More, D.S.; Fernandes, B.G. Novel three phase flux reversal machine with full pitch winding. In Proceedings
of the 7th International Conference on Power Electronics, Daegu, Korea, 22–26 October 2007; pp. 1007–1012.
27. Yang, H.; Zhu, Z.Q.; Lin, H.; Li, H.Y.; Lyu, S. Analysis of consequent-pole flux reversal permanent magnet
machine with biased flux modulation theory. IEEE Trans. Ind. Electron. 2020, 67, 2107–2121. [CrossRef]
28. Luo, J.; Kou, B.; Zhang, H.; Qu, R. Development of a consequent pole transverse flux permanent magnet
linear machine with passive secondary structure. CES Trans. Electr. Mach. Syst. 2019, 3, 39–44. [CrossRef]
29. Zhu, X.; Hua, W. Back-EMF waveform optimization of flux-reversal permanent magnet machines. AIP Adv.
2017, 7, 1–5. [CrossRef]
30. Dmitrievskii, V.; Prakht, V.; Sarapulov, S.; Askerov, D. A multipole single-phase SMC flux reversal
motor for fans. In Proceedings of the 2016 XXII International Conference on Electrical Machines (ICEM),
Lausanne, Switzerland, 4–7 September 2016; pp. 53–59.
31. Dmitrievskii, V.; Prakht, V.; Pozdeev, A.; Klimarev, V.; Mikhalitsyn, A. Angular grinder with new flux reversal
motor. In Proceedings of the 2015 International Conference on Electrical Machines and Systems (ICEMS),
Pattaya, Thailand, 25–28 October 2015; pp. 1366–1371.
32. Dmitrievskii, V.; Prakht, V. Gearless generator with magnets on the stator for wind turbine. In Proceedings
of the Wind Europe Conference 2018, Hamburg, Germany, 25–28 September 2018; pp. 1–8.
33. Vidhya, B.; Srinivas, K.N. Effect of stator PM thickness and rotor geometry modifications on the minimization
of cogging torque of flux reversal machine. Turk. J. Electr. Eng. Comput. Sci. 2017, 25, 4907–4922. [CrossRef]
34. Heoung, T.K.; Sung, H.W.; Bong, K.; Lee, J. Reduction of cogging torque in flux reversal machine by rotor
teeth pairing. In Proceedings of the IEEE International Magnetics Conference (INTERMAG), Nagoya, Japan,
4–8 April 2005; pp. 3964–3966.
35. Gyeong, C.L.; Jung, T.U. Cogging torque reduction design of dual stator radial flux permanent magnet
generator for small wing turbine. In Proceedings of the IEEE 2013 Tencon-Spring, Sydney, NSW, Australia,
17–19 April 2013; pp. 85–89.
36. Xu, W.; Zhu, J.; Zhang, Y.; Hu, J. Cogging torque reduction for radially laminated flux-switching permanent
magnet machine with 12/14 poles. In Proceedings of the IECON 2011-37th Annual conference on IEEE
Industrial Electronics Society, Melbourne, VIC, Australia, 7–10 November 2011; pp. 3590–3595.
37. Yusivar, F.; Roy, V.H.S.; Gunawan, R.; Halim, A. Cogging torque reduction with pole slot combination and
notch. In Proceedings of the IEEE International Conference on Electrical Engineering and Computer Science,
Kuta, Indonesia, 24–25 November 2014; pp. 260–263.

Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional
affiliations.

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

78
energies
Article
A Study on the Rotor Design of Line Start
Synchronous Reluctance Motor for IE4 Efficiency
and Improving Power Factor
Hyunwoo Kim 1 , Yeji Park 1 , Seung-Taek Oh 1 , Hyungkwan Jang 1 , Sung-Hong Won 2 ,
Yon-Do Chun 3 and Ju Lee 1, *
1 Department of Electrical Engineering, Hanyang University, Seoul 04763, Korea;
[email protected] (H.K.); [email protected] (Y.P.); [email protected] (S.-T.O.);
[email protected] (H.J.)
2 Department of Electrical Engineering, Dongyang Mirae University, Seoul 08221, Korea;
[email protected]
3 Electric Machines and Drives Research Center, Korea Electrotechnology Research Institute,
Changwon 51543, Korea; [email protected]
* Correspondence: [email protected]; Tel.: +82-2220-0342

Received: 15 October 2020; Accepted: 3 November 2020; Published: 4 November 2020

Abstract: As international regulations of motor efficiency are strengthened, the line-start synchronous
reluctance motor (LS-SynRM) is being studied to improve the efficiency of the electrical motor in
industrial applications. However, in industrial applications, the power factor is also an important
performance index, but the LS-SynRM has poor power factor due to the saliency characteristic.
In this paper, the rotor design of LS-SynRM is performed to improve the efficiency and power factor.
First, the barrier design is performed to improve the efficiency and power factor using the response
surface method (RSM). Second, the rotor slot design is performed according to the length of bar for
synchronization. Lastly, the rib design is performed to satisfy the power factor and the mechanical
reliability. The final model through the design process is analyzed using finite element analysis (FEA),
and the objective performance is satisfied. To verify the FEA result, the final model is manufactured,
and experiment is performed.

Keywords: finite element analysis; international electrotechnical committee; line start synchronous
reluctance motor; power factor; super premium efficiency

1. Introduction
In industrial applications, electrical energy consumption of motors account for 35% to 40%
of electrical energy generated in the world. If this electrical energy consumption can be reduced,
several environmental impacts, such as the emissions of CO2 or global warming, can be reduced [1,2].
For this reason, international regulations regarding motor efficiency were enacted by the International
Electrotechnical Committee (IEC) 60034-30. According to IEC 60034-30, motor efficiency was classified
as IE1 to IE5. Recently, the efficiency standards of industrial applications have been strengthened, and
industrial motors may need to meet IE4 or IE5 efficiency class [3].
With various types of motors, the three-phase squirrel-cage induction motor (SCIM) is the
most used, because of its simple structure, line-start ability, robustness, and low manufacturing cost.
However, it is evident that the induction motor (IM) does not meet high efficiency due to rotor copper
loss [4–6]. Due to a global trend on improving the efficiency of electric motors, there is ongoing
research that focuses on this area. In References [7,8], the aluminum in the rotor slot is replaced by
copper to reduce the rotor copper loss. Copper die-casting induction motors can improve efficiency by

Energies 2020, 13, 5774; doi:10.3390/en13215774 www.mdpi.com/journal/energies

79
Energies 2020, 13, 5774

about 2–3% compared to the aluminum die-casting induction motors. However, copper has a higher
melting point than aluminum, so a die-casting process is difficult, and the manufacturing cost increases.
In industrial variable-speed applications, a variable-speed drive (VSD) is used to improve the efficiency
of motors [9–11]. When using the VSD system, a synchronous motor can be controlled to improve the
efficiency. However, the synchronous motor along with the VSD system is more expensive than SCIMs.
Line-start synchronous machines (LSSMs) are actively studied because they can meet the efficiency
of IE4. In addition, a LSSM is direct-on-line (DOL) motor, such as IM, so it does not require an
inverter, in which the cost of electric system can be reduced [12–17]. LSSM can be classified into two
types; line-start permanent magnet synchronous motor (LSPM) and line-start synchronous reluctance
motor (LS-SynRM). LSPM has the high efficiency and power factor because of the permanent magnet.
However, the permanent magnet of LSPM must use a rare-earth magnet, such as neodymium, because of
the high starting current. Furthermore, the resources of rare-earth magnets are limited, and lead to
high costs [18,19]. Therefore, the LS-SynRM has received attention for its high efficiency and for the
fact that it does not use a rare-earth magnet.
A reactive power of an electrical motor depends on the power factor. If the reactive power is
increased by the low power factor, the copper loss in stator winding increases under the same output
power and input voltage condition because of the current increases [20]. This increased copper loss
decreases the efficiency of the electrical motor. For this reason, there are the power factor standards of
industrial motors, according to IEC 60034. The electrical motor must meet the power factor standards
to use industrial applications [21]. However, LS-SynRM has a lower power factor than IM and LSPM
because of a saliency characteristic and may not meet the power factor standards [22–24]. Therefore,
the design method considering the power factor is required.
This paper is a design process of LS-SynRM, for super premium efficiency, and for improving the
power factor. The efficiency and the power factor of LS-SynRM is determined by dq-axis inductance,
which is determined by the barrier design. Therefore, the design parameters are selected for IE4
efficiency and the power factor. Based on the selected design parameters, the barrier optimal design is
performed using the response surface method (RSM) and finite element analysis (FEA). In addition,
the rotor slot design is performed for synchronization and the rib design is performed to improve the
power factor and satisfy the mechanical reliability through the safety factor. To verify the efficiency
and the power factor of the design result, the final model is manufactured, and the experiment is
performed to verify the IE4 efficiency and power factor.
This paper is organized as follows. In Section 2, the characteristics of LS-SynRM are discussed
in respect to the structure and the operation in asynchronous and synchronous speed. In Section 3,
the characteristic of the reference machine is analyzed. In Section 4, the design of LS-SynRM is
performed using RSM and FEA and the final model is analyzed using FEA. In Section 5, to verify FEA
results, the experiment is performed and the results of LS-SynRM are compared with FEA results.
Finally, Section 6 contains the conclusion.

2. Characteristic of LS-SynRM

2.1. Structure of LS-SynRM


Figure 1 shows the rotor structure of a general LS-SynRM. LS-SynRM has a structure in which the
squirrel cage slots are placed on the outer side of the rotor, and the barrier is arranged on the inner side
of the rotor. Because of this structure, LS-SynRM operates as IMs at the asynchronous speed and as
SynRM at the synchronous speed. Moreover, the rotor design of LS-SynRM is classified as a rotor slot
and barrier. The rotor slot design determines a starting characteristic of LS-SynRM and the barrier
design determines the performance of LS-SynRM, such as the efficiency and power factor. Therefore,
the rotor slot design is important to reach the synchronous speed, and the barrier design is important
to satisfy the super-premium efficiency (IE4).

80
Energies 2020, 13, 5774

Figure 1. Rotor structure of the line-start synchronous reluctance motor (LS-SynRM).

2.2. Asynchronous Speed


In asynchronous speed, the LS-SynRM has the slip, which generates an induced voltage on a
conductor in a rotor slot. An induced current that is generated by the induced voltage provides
the magnetic torque that LS-SynRM can be synchronization. Moreover, in case of the LS-SynRM,
the reluctance torque is generated due to a saliency difference, which is the difference between dq-axis
inductance. Therefore, the torque of LS-SynRM is expressed as Equation (1) [25,26].

Te = Tcage + Trel sin(2sωt + α) (1)

where Te is the torque of LS-SynRM at asynchronous speed, Tcage is the magnetic torque generated by
the induced current, Trel is the amplitude of reluctance torque produced by the saliency, s is slip, ω is
the electrical angular frequency, and α is the phase angle of pulsating torque.
The magnetic torque is constant and depends on the slip as in the induction motor, and the
average of reluctance torque is zero, but this torque pulsates twice the slip frequency. Figure 2 shows a
speed-torque curve of LS-SynRM at an asynchronous speed. The amplitude of the pulsating torque
and the average torque depend on the slip. In a synchronous speed, the maximum torque is called the
pull-out torque, which is the maximum load torque that can be synchronized [26].

Figure 2. Speed-torque curve of LS-SynRM.

2.3. Synchronous Speed


At a synchronous speed, there are no eddy currents in the rotor slot because the slip is zero.
Therefore, the LS-SynRM is operated as the synchronous reluctance motor and the efficiency is

81
Energies 2020, 13, 5774

improved compared to SCIM due to the secondary copper loss. The efficiency of LS-SynRM is
expressed as follows.
e ωe
η = PPout = PoutP+outP = Te ωe +PTcopper +Pcore
in loss (2)
Te = 32 P2 (Ld − Lq )Ia2 sin 2β
where η is the efficiency, Pout is mechanical power, Pin is electrical input power, Ploss is total loss of
motor, Te is torque of motor, ωe is synchronous speed, Pcopper is copper loss, and Pcore is core loss, P is
the number of pole, Ld and Lq are d-, q-axis inductance, respectively, Ia is the current, and β is the
current phase angle.
The efficiency is determined by the reluctance torque, the core loss, and the copper loss. The core
loss depends on the magnetic flux density and input frequency, which is determined when designing
electric machines. Therefore, the copper loss must be reduced to improve the efficiency. This should
increase the torque per current, which should increase the rotor saliency difference.
The power factor of the LS-SynRM is determined by the phase difference between the voltage and
current. Figure 3a shows a vector diagram of the LS-SynRM, and the power factor of LS-SynRM can be
expressed as follows.
ρ−1 L
cos ϕ =  , ρ= d (3)
ρ2 Lq
cos2 β
+ 1
2 sin β

where cosϕ is the power factor, and ρ is the saliency ratio.

(a) (b)

Figure 3. Characteristic of LS-SynRM (a) vector diagram, and (b) power factor.

The power factor is determined by the saliency ratio and the current phase angle. Figure 3b
shows the power factor according to the saliency ratio and the current phase angle. The power factor
increases as the saliency ratio increases. Therefore, the saliency difference and ratio are important
design parameters for IE4 efficiency and improving the power factor.

3. Reference Machine
In order to compare the efficiency and power factor of LS-SynRM, 2.2 kW four pole IM is selected
as the reference motor. Figure 4 show the FEA model and Table 1 shows the specification of reference
machine. ANSYS Maxwell was used to analyze the performance of the reference machine.

82
Energies 2020, 13, 5774

Figure 4. Finite element analysis (FEA) model of reference machine.

Table 1. Specification of reference machine.

Item Value Unit


Output power 2.2 kW
Line voltage 380 Vrms
Input frequency 60 Hz
Outer diameter (stator/rotor) 180/109.4 mm
Inner diameter (stator/rotor) 110/32 mm
Number of slot (stator/rotor) 36/28 –
Stack length 120 mm
Air gap length 0.3 mm

Table 2 shows the FEA and experiment results of the reference machine. The efficiency of the
reference machine is 90.7% that is IE3 efficiency according to IEC 60034. Because IE4 efficiency is 91%,
the reference machine does not satisfy IE4 efficiency. In general, the induction motor has the rotor
copper loss due to the slip, and this loss accounts for 23% of total losses in Table 2. If the rotor copper
loss is reduced, the efficiency of the electrical machine can satisfy IE4 efficiency. LS-SynRM does not
have the rotor copper loss because this motor is operated at a synchronous speed, where slip is zero.
Therefore, LS-SynRM can satisfy IE4 efficiency, and the design process of LS-SynRM has studied to
improve the efficiency.

Table 2. FEA and experiment result of reference model.

Value
Item Unit
FEA Experiment
Power 2.2 2.2 kW
Speed 1764.96 1762 rpm
Torque 12 11.9 Nm
Core loss 58.27 63 W
Stator copper loss 72.6 78 W
Rotor copper loss 40.53 51 W
Total loss 214.3 220 W
Efficiency 91.2 90.7 %
Power factor 83.8 80 %

83
Energies 2020, 13, 5774

4. Design of LS-SynRM

4.1. Design Process


Figure 5 shows the design process of LS-SynRM for IE4 efficiency and improving the power factor.
The stator, rotor, and number of rotor slot are constrained under the same conditions as the reference
machine. The main design process is three steps. First, the design of the barrier is performed to
improve the efficiency and power factor using RSM. The efficiency and power factor of LS-SynRM
are determined by the saliency difference and ratio. Therefore, there are two steps in the design of
the barriers: the number of barriers and thickness of barriers and segments. Second, the design of
the rotor slot is performed to reach LS-SynRM into the synchronous speed. Because synchronization
is determined by the resistance and leakage inductance of the rotor bar, the design is performed
according to the depth of the rotor slot. Lastly, the design of each rib is performed considering the
power factor and mechanical reliability. The thickness of the ribs affects the leakage flux, which affect
the performance of LS-SynRM, and is determined by the thickness of each rib. Therefore, the thickness
of each rib must be designed considering the performance and safety factor. If the efficiency and
power factor does not satisfy, the design parameter and range are reselected, and RSM is performed to
satisfy the performance. Table 3 shows IEC 60034 standard and the design objective of the efficiency
and power factor. Considering the margin based on Table 2, FEM performance is selected as 91.5%
efficiency and 81% power factor.

Figure 5. Design process of LS-SynRM.

84
Energies 2020, 13, 5774

Table 3. International Electrotechnical Committee (IEC) 60034 standard and design objective of the
efficiency and power factor.

Items IEC 60034 Design Objective Unit


Efficiency 91 ≥91.5 %
Performance
Power factor 77 ≥81 %

4.2. Design of Barrier


From Equation (3), the power factor of LS-SynRM is determined by the saliency ratio, which is
ratio d-axis inductance to q-axis inductance. The dq-axis inductances are determined by the number of
barriers and thickness of flux barrier and segments [27]. Therefore, the barrier design is performed
according to the number of barriers and thickness of flux barrier and segments.

4.2.1. Number of Barrier


Figure 6 shows the FEA model according to the number of barriers. The number of rotor slots is the
same as the reference machine, and the total length of barrier and segment is constant, for comparison
under the same conditions. Furthermore, for each layer of each model, the thickness of the barrier and
segment are compared equally. Table 4 shows the comparison of the dq-axis inductances, the saliency
ratio, the power factor, and efficiency under the rated current condition. From Table 4, the larger the
saliency ratio, the larger the power factor. Moreover, the efficiency of each model satisfies more than
IE4 efficiency. Therefore, the design of the barrier, according to the thickness of the flux barrier and
segments, is performed based on model 3.

(a) (b) (c)

Figure 6. FEA model according to the number of barriers (a) model 1, (b) model 2, and (c) model 3.

Table 4. FEA result of LS-SynRM according to the number of barriers.

Ld Lq Saliency Power Factor


Model
(mH) (mH) Ratio (%)
Model 1 33.6 301.6 8.97 78.7
Model 2 31.7 297.7 9.38 79.3
Model 3 31.2 309.5 9.92 80.5

4.2.2. Thickness of Flux Barriers and Segments


To maximize the power factor, the optimal design is performed using RSM and FEA. Figure 7
shows the design parameters for the optimal design [28]. The range of design parameters is selected
within the rotor constraints, as shown in Table 5. Furthermore, the objective function is maximizing
the power factor and the efficiency. Figure 8a shows the RSM result and Figure 8b shows the optimal

85
Energies 2020, 13, 5774

design result. Table 6 shows the optimal design result. In Table 6, RSM and FEA results are similar,
so the optimal result is valid.

Figure 7. Design parameters for the optimal design using the response surface method (RSM).

Table 5. Range of design parameters.

Items αr (degM) Kw Wshaft (mm)


Range 4.5–9 0.4–0.6 3–6

(a) (b)

Figure 8. Optimization using RSM under rated current condition. (a) RSM result, (b) optimal
design result.

Table 6. Optimal design result.

Power Factor Efficiency


Parameter
(%) (%)
αr Wshaft
Kw RSM FEA RSM FEA
(degM) (mm)
7.236 0.5609 4 81.19 81.02 91.59 91.75

4.3. Design of Rotor Slot


In general, the starting characteristic is determined by the rotor resistance and leakage inductance.
This rotor resistance and leakage is determined by the depth and thickness of the rotor slot. However,
the barriers are constraints on the thickness and depth of the rotor slot. Therefore, for synchronization,

86
Energies 2020, 13, 5774

the design parameters of the rotor slot are determined, as shown in Figure 9a. In order to analyze
synchronization, the analysis, according the selected design parameters, is performed by the time-step
FEA, which is the mechanical and electromagnetic transient analysis [29,30], considering inertia of
LS-SynRM. Figure 9b shows the time step FEA result. The synchronization region is determined by
the depth of the rotor slot. Figure 10 shows the time step FEA result for the three selected points.
If synchronization fails, the torque and speed pulsate. As a result, the design parameter is determined
as Tbar1 = 20 (mm) and Tbar2 = 13 (mm).

Tbar 

Tbar

(a) (b)

Figure 9. Design parameter of rotor slot and FEA result (a) design parameter, (b) FEA result.

(a) (b)

Figure 10. Speed and torque curve of the selected three points (a) speed curve (b) torque curve.

4.4. Design of Ribs


When LS-SynRM rotates, the centrifugal force is focused on the outer rib. If the thickness of the
outer rib is designed to be small, the LS-SynRM is mechanically damaged during operation. Therefore,
the design of rib must be performed for use as the industrial application. However, the thickness of
the outer rib also affects the power factor, so the design of the rib is performed considering the power
factor and mechanical reliability. The mechanical reliability is determined by a safety factor, as shown
in the following equation. Generally, the mechanical reliability is ensured if the safety factor is higher
than 1.5 [31].
σyield
SF = (4)
σmax

87
Energies 2020, 13, 5774

where SF is the safety factor, σyield is the tensile yield strength of material, σmax is the stress of material
during rotation.
Because of the structure of LS-SynRM, there are several bridges that reduce the mechanical stress
at the rib. Therefore, the design parameters are selected, as shown in Figure 11. Considering the
manufacture, the design parameters are designed to be 0.3 (mm) or more. Figure 12 shows the power
factor and the safety factor, according to the design parameters. As the thickness of each rib increases,
the power factor is reduced because the leakage magnetic flux is increased. Furthermore, the safety
factor is increased due to reduced mechanical stress. Considering the objective performance in Table 3,
the design parameters are designed as Trib1 = 0.3 (mm) and Trib2 = 0.7 (mm).

Figure 11. Design parameter for the power factor and safety factor.

(a) (b)

Figure 12. FEA result according to the design parameter of rib (a) safety factor, and (b) power factor.

4.5. Design Result


Figure 13 shows the design result through the design process of LS-SynRM. Table 7 shows the
FEA result of LS-SynRM. The efficiency of LS-SynRM is 91.7% and the power factor of LS-SynRM is
81.2%. Compared with Table 2, the efficiency is improved about 0.5%, but the power factor is decreased
about 2.6% than the reference machine. The efficiency satisfies the IE4 efficiency and the power factor
satisfies the IEC 60034 standard for the industrial application. Therefore, the LS-SynRM can improve
the efficiency compared with IM and be used as an industrial electric machine.

88
Energies 2020, 13, 5774

Figure 13. Final model of LS-SynRM.

Table 7. FEA result of the final LS-SynRM.

Item Value Unit


Power 2.2 kW
Speed 1800 rpm
Torque 11.7 Nm
Core loss 53.3 W
Stator copper loss 87.78 W
Rotor copper loss 12.3 W
Total loss 201.4 W
Efficiency 91.7 %
Power factor 81.2 %
Safety factor 1.56 -

5. Experimental Validation

5.1. Manufacture of LS-SynRM


To verify the FEA result, the designed LS-SynRM is manufactured. Figure 14a shows the rotor
core of the manufactured LS-SynRM. A die-casting process is performed to inject the melted aluminum
into the rotor slot. During the die-casting process, a high press is applied to the rotor slot so that the
melted aluminum is injected in the rotor slot. Figure 14b shows the endplate, which is required to
prevent the melted aluminum from being injected into the barrier. Figure 14c shows the cross-section
of the rotor and the aluminum is filled in the rotor slot. Figure 14d shows the rotor with the end-ring.
The width of the end-ring is the same as the reference machine and the height of the end-ring is the
difference between the deepest rotor slot depth and the outer diameter of the rotor.

5.2. Experiment Result


The experiment is performed to verify the FEA result. Figure 15 shows the experiment environment.
The dynamometer motor is induction motor as 2.2 kW load. Using voltage and frequency control,
LS-SynRM runs up and reaches a synchronous speed. After the LS-SynRM reaches a synchronous
speed, the load torque is applied through the dynamometer motor to measure the efficiency and
power factor of LS-SynRM. In the power analyzer, the efficiency and power factor are calculated
using measured voltage, current, torque, and speed. Table 8 shows the comparison of FEA and the
experiment result of LS-SynRM. When comparing the experiment and FEA results, the experiment
current increases by about 5% than the FEA current. This reduces the power factor and increases
the copper loss, as shown in Table 8. The power factor of FEA and experiment is 81.2% and 77.3%,
respectively. However, the efficiencies of FEA and the experiment are similar because the total losses
are similar. As a result, it is confirmed that the IE4 class efficiency and the power factor in IEC 60034
is satisfied.

89
Energies 2020, 13, 5774

(a) (b)

(c) (d)

Figure 14. Manufactured LS-SynRM (a) rotor core, (b) endplate for die-casting, (c) cross-section of
rotor, (d) rotor with the end-ring.

Figure 15. The experiment environment and test dynamometer.

90
Energies 2020, 13, 5774

Table 8. Comparison of FEA and experiment result.

Value
Item Unit
FEA Experiment
Power 2.2 2.2 kW
Torque 11.7 11.6 Nm
Current 4.56 4.8 Arms
Core loss 53.3 45.9 W
Stator copper loss 87.78 100 W
Rotor copper loss 12.3 0 W
Total loss 201.4 201.7 W
Efficiency 91.7 91.6 %
Power factor 81.2 77.3 %

6. Conclusions
In this paper, the design method of LS-SynRM is studied to satisfy the efficiency and power factor
of IEC 60034 standard. In addition, LS-SynRM is designed considering the mechanical reliability for use
in the industrial application. The performance of LS-SynRM is analyzed according to several design
parameters, such as the number of barriers, the thickness of the barriers, and the segments, the depth of
the rotor slot, and thickness of each rib. The design of the barrier is performed to satisfy the performance
of LS-SynRM using RSM. The design of the rotor slot is performed, and the synchronization region is
determined by the depth of the rotor slot. The design of each rib is performed, and the thickness of
each rib is designed considering the power factor and the safety factor. Considering the efficiency,
the power factor, and the safety factor, the optimal LS-SynRM is designed. The final LS-SynRM satisfies
the efficiency of 91.7%, the power factor of 81.2%, and the safety factor 1.56. The final LS-SynRM is
manufactured and the experiment is performed to verify the FEA result. As a result, the efficiency and
power factor satisfy the IEC 60034 standard.

Author Contributions: Conceptualization, H.K.; methodology, H.J.; validation, S.-T.O.; formal analysis, H.K.
and S.-T.O.; investigation, H.K. and Y.P.; resources, S.-H.W.; data curation, Y.P.; writing—original draft preparation,
H.K.; writing—review and editing, S.-T.O. and H.J.; supervision, S.-H.W. and J.L.; project administration,
J.L.; funding acquisition, Y.-D.C. and S.-H.W. All authors have read and agreed to the published version of
the manuscript.
Funding: This research was funded by the Korea Agency for Infrastructure Technology Advancement (KAIA) grant
funded by the Korea government Ministry of Ministry of Land, Infrastructure and Transport (20CTAP-C157784-01)
and in part by a Korea Institute of Energy Technology Evaluation and Planning grant funded by the Korea
government (20192010106780, A Construction and Operation of Open Platform for Next-Generation Super
Premium Efficiency (IE4) Motors).
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Almeida, A.T.D.; Ferreira, F.J.T.E.; Baoming, G. Beyond Induction Motors-Technology Trends to Move Up
Efficiency. IEEE Ind. Appl. 2014, 50, 2103–2114.
2. Almeida, A.T.D.; Ferreira, F.J.T.E.; Duarte, A.Q. Technical and Economical Considerations on Super
High-Efficiency Three-Phase Motors. IEEE Ind. Appl. 2014, 50, 1274–1285.
3. Ferreira, F.J.T.E.; Baoming, G.; Almeida, A.T.D. Reliability and Operation of High-Efficiency Induction
Motors. IEEE Ind. Appl. 2016, 52, 4628–4637.
4. Rafajdus, P.; Hrabovcova, V.; Lehocky, P.; Makys, P.; Holub, F. Effect of Saturation on Field Oriented Control
of the New Designed Reluctance Synchronous Motor. Energies 2018, 11, 3223. [CrossRef]
5. Aguba, V.; Muteba, M.; Nicolae, D.V. Transient Analysis of a Start-up Synchronous Reluctance Motor
with Symmetrical Distributed Rotor Cage Bars. In Proceedings of the 2017 IEEE AFRICON, Cape Town,
South Africa, 18–20 September 2017.
6. Xie, Y.; Pi, C.; Li, Z. Study on Design and Vibration Reduction Optimization of High Starting Torque Induction
Motor. Energies 2019, 12, 1263. [CrossRef]

91
Energies 2020, 13, 5774

7. Kim, D.-J.; Hong, D.-K.; Choi, J.-H.; Chun, Y.-D.; Woo, B.-C.; Koo, D.-H. An Analytical Approach for
a High Speed and High Efficiency Induction Motor Considering Magnetic and Mechanical Problems.
IEEE Trans. Magn. 2013, 49, 2319–2322. [CrossRef]
8. Zhang, Q.; Liu, H.; Zhang, Z.; Song, T. A Cast Copper Rotor Induction Motor for Small Commercial EV
Traction: Electromagnetic Design, Analysis, and Experimental Tests. CES Trans. Electr. Mach. Syst. 2018,
2, 417–424. [CrossRef]
9. Kazakbaev, V.; Prakht, V.; Dmitrievskii, V.; Ibrahim, M.N.; Oshurbekov, S.; Sarapulov, S. Efficiency Analysis of
Low Electric Power Drives Employing Induction and Synchronous Reluctance Motors in Pump Applications.
Energies 2019, 12, 1144. [CrossRef]
10. Aarniovuori, L.; Kolehmainen, J.; Kosonen, A.; Niemela, M.; Chen, H.; Cao, W.; Pyrhonen, J. Application
of Calorimetric Method for Loss Measurement of a SynRM Drive System. IEEE Trans. Ind. Electron. 2016,
64, 2005–2015. [CrossRef]
11. Goetzler, W.; Sutherland, T.; Reis, C. Energy Savings Potential and Opportunities for High Efficiency Electric Motors
in Residential and Commercial Equipment; United States Department of Energy: Washington, DC, USA, 2013.
12. Ghoroghchian, F.; Aliabad, A.D.; Amiri, E. Design improvement of dual-pole LSPM synchronous motor.
IET Electr. Power Appl. 2019, 13, 742–749. [CrossRef]
13. Ghahfarokhi, M.M.; Aliabad, A.D.; Boroujeni, S.T.; Amiri, E.; Faradonbeh, V.Z. Analytical modelling and
optimization of line start LSPM synchronous motors. IET Electr. Power Appl. 2020, 14, 398–408.
14. Mingardi, D.; Bianchi, N. Line-Start PM-Assisted Synchronous Motor Design, Optimization, and Tests.
IEEE Trans. Ind. Electron. 2017, 64, 9739–9747. [CrossRef]
15. Liu, H.C.; Lee, J. Optimum Design of an IE4 Line-Start Synchronous Reluctance Motor Considering
Manufacturing Process Loss Effect. IEEE Trans. Ind. Electron. 2018, 65, 3104–3114. [CrossRef]
16. Tampio, J.; Kansakangas, T.; Suuriniemi, S.; Kolehmainen, J.; Kettunen, L.; Ikaheimo, J. Analysis of
Direct-On-Line Synchronous Reluctance Machine Start-Up Using a Magnetic Field Decomposition. IEEE Trans.
Ind. Appl. 2017, 53, 1852–1859.
17. Liu, H.-C.; Hong, H.-S.; Cho, S.; Lee, J.; Jin, C.-S. Bubbles and Blisters Impact on Diecasting Cage to the
Designs and Operations of Line-Start Synchronous Reluctance Motors. IEEE Trans. Magn. 2017, 53, 8202504.
[CrossRef]
18. Dent, P. Rare earth elements and permanent magnets. J. Appl. Phys. 2012, 111, 07A721. [CrossRef]
19. Goss, J.; Popescu, M.; Staton, D. A comparison of an interior permanent magnet and copper rotor induction
motor in a hybrid electric vehicle application. In Proceedings of the International Electric Machines & Drives
Conference, Chicago, IL, USA, 12–15 May 2013; pp. 220–225.
20. Kazakbaev, V.; Parkht, V.; Dmitrievskii, V.; Oshurbekov, S.; Golovanov, D. Life Cycle Energy Cost Assessment
for Pump Units with Various Types of Line-Start Operating Motors Including Cable Losses. Energies 2020,
13, 3546–3559.
21. Kim, H.; Park, Y.; Liu, H.-C.; Han, P.-W.; Lee, J. Study on Line-Start Permanent Magnet Assistance Synchronous
Reluctance Motor for Improving Efficiency and Power Factor. Energies 2020, 13, 384. [CrossRef]
22. Joo, K.-J.; Kim, I.-G.; Lee, J.; Go, S.-C. Robust Speed Sensorless Control to Estimated Error for PMa-SynRM.
IEEE Trans. Magn. 2017, 53, 8102604.
23. Ozcelik, N.G.; Dogru, U.E.; Imeryuz, M.; Ergene, L.T. Synchronous Reluctance Motor vs. Induction Motor at
Low-Power Industrial Applications: Design and Comparison. Energies 2019, 12, 2190. [CrossRef]
24. Ogunjuyigbe, A.S.O.; Jimoh, A.A.; Ncolae, D.V.; Obe, E.S. Analysis of synchronous reluctance machine with
magnetically coupled three-phase windings and reactive power compensation. IET Electr. Power Appl. 2010,
4, 291–303. [CrossRef]
25. Kersten, A.; Liu, Y.; Pehrman, D.; Thiringer, T. Rotor Design of Line-Start Synchronous Reluctance Machine
With Round Bars. IEEE Ind. Appl. 2019, 55, 3685–3696. [CrossRef]
26. Gamba, M.; Pellegrino, G.; Vagati, A.; Villata, F. Design of a Line-Start Synchronous Reluctance Motor.
In Proceedings of the 2013 International Electric Machines & Drives Conference, Chicago, IL, USA,
12–15 May 2013.
27. Pellegrino, G.; Cupertino, F.; Gerada, C. Automatic Design of Synchronous Reluctance Motors Focusing on
Barrier Shape Optimization. IEEE Trans. Ind. Appl. 2015, 51, 1465–1474. [CrossRef]
28. Kim, K.-C.; Ahn, J.S.; Won, S.H.; Hong, J.-P.; Lee, J. A Study on the Optimal Design of SynRM for the High
Torque and Power Factor. IEEE Trans. Magn. 2007, 43, 2543–2545. [CrossRef]

92
Energies 2020, 13, 5774

29. Liu, H.C.; Seol, H.S.; Kim, J.Y.; Lee, J. Design and Analysis of an IE4 Class Line-Start Synchronous Reluctance
Motor Considering Total Loss and Starting Performance. J. Electron. Mater. 2019, 48, 1386–1394. [CrossRef]
30. Wang, Y.; Chau, K.T.; Chan, C.C.; Jiang, J.Z. Transient Analysis of a New Outer-Rotor Permanent-Magnet
Brushless DC Drive Using Circuit-Field-Torque Coupled Time-Stepping Finite-Element Method.
IEEE Trans. Magn. 2002, 38, 1297–1300. [CrossRef]
31. Lee, J.-K.; Jung, D.-H.; Lim, J.; Lee, K.-D.; Lee, J. A Study on the Synchronous Reluctance Motor Design for
High Torque by Using RSM. IEEE Trans. Magn. 2018, 54, 8103005. [CrossRef]

Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional
affiliations.

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

93
energies
Article
Investigation of Torque Performance and Flux Reversal
Reduction of a Three-Phase 12/8 Switched Reluctance Motor
Based on Winding Arrangement
Ruchao Pupadubsin, Seubsuang Kachapornkul, Prapon Jitkreeyarn, Pakasit Somsiri and Kanokvate Tungpimolrut *

Machines and Power Conversions Research Team, National Electronics and Computer Technology Center,
National Science Technology Development Agency, Pathum Thani 12120, Thailand;
[email protected] (R.P.); [email protected] (S.K.);
[email protected] (P.J.); [email protected] (P.S.)
* Correspondence: [email protected]

Abstract: The goal of this paper is to present a comparative analysis of two types of winding arrange-
ments for a three-phase 12/8 switched reluctance motor (SRM), where short- and fully-pitched winding
arrangements under unipolar operation are considered. From the analytical results, the short-pitched
winding has the best torque per copper weight ratio. The core loss based on counting the number of flux
reversals in the stator yoke for each winding arrangement is also proposed and mentioned. To reduce the
magnetic flux reversals in the stator core, changing the direction of the magnetic flux path by modifying
the winding polarities of the short-pitched winding could reduce 10–13% of core loss compared to the
conventional winding. A 1 kW, 12/8 SRM prototype for the ventilation fan application is constructed
and tested in order to verify the design consideration of winding configuration. At the rated condition, a
Citation: Pupadubsin, R.; maximum efficiency around 82.5% could be achieved.
Kachapornkul, S.; Jitkreeyarn, P.;
Somsiri, P.; Tungpimolrut, K. Keywords: switched reluctance machine; winding arrangement; static torque; dynamic torque;
Investigation of Torque Performance
torque ripple; core loss; flux reversal; leakage-flux; magnetic flux path
and Flux Reversal Reduction of a
Three-Phase 12/8 Switched
Reluctance Motor Based on Winding
Arrangement. Energies 2022, 15, 284.
1. Introduction
https://doi.org/10.3390/en15010284
The switched reluctance motor (SRM) [1,2] is a double salient machine having a
Academic Editors: Vladimir Prakht,
simple magnetic structure. It has only concentrated windings on the stator and without
Mohamed N. Ibrahim and Vadim
any coil winding or permanent magnet on the rotor, resulting in cost savings for the
Kazakbaev
manufacturing process. Other advantages of SRM include high starting torque, fault
Received: 29 November 2021 tolerant, high efficiency over wide speed range and broad choice of converters depending
Accepted: 29 December 2021 on the application requirements; these make SRM an attractive alternative to conventional
Published: 1 January 2022 drive in numerous applications.
Publisher’s Note: MDPI stays neutral
Many modern applications require higher efficiency and higher performance machines
with regard to jurisdictional claims in
with reducing production cost. The challenge issue of SRM design techniques is to increase
published maps and institutional affil- torque density while minimizing torque ripple, as these are major disadvantages of SRMs
iations. when compared to permanent–magnet (PM) motors. Most of the design methods are
directly implemented via the design of motor geometries such as stator/rotor pole arc,
stator/rotor pole shape, air gap, and stack length. To reduce the manufacturing cost,
research work in [3] introduced the optimization algorithm for SRM design to achieve
Copyright: © 2022 by the authors. a maximum torque at the minimum mass of the motor. The recent optimization design
Licensee MDPI, Basel, Switzerland. technique for finding the maximum torque density of 12/8 SRM was revealed in [4]. This
This article is an open access article algorithm, based on the Grey Wolf Optimizer (GWO), can increase torque by 120% at the
distributed under the terms and same outer value as the original design.
conditions of the Creative Commons Changing winding arrangements is a simple method that does not require the machine
Attribution (CC BY) license (https://
structure to be redesigned, and a substantial increase in efficiency or performance can be
creativecommons.org/licenses/by/
achieved. For example, with a fully-pitched winding arrangement, as reported in [5–7],
4.0/).

Energies 2022, 15, 284. https://doi.org/10.3390/en15010284 95 https://www.mdpi.com/journal/energies


Energies 2022, 15, 284

torque capability can be increased by over 30%, resulting in higher efficiency than the
conventional winding arrangement. The fully-pitched winding, which can operate with
either unipolar or bipolar operation, uses the mutual-coupling among the phases for torque
production and enhancement of torque density. In [8,9], an analysis, design, and SRM
operation of the fractionally-pitched winding arrangement that also utilizes the mutual–
coupling between phases are discussed. However, the fractionally-pitched winding is
suitable for bipolar operation rather than unipolar operation. Another investigation on
winding arrangements [10,11] uses finite element analysis to study short- and long-flux
paths using single-phase and double-phase excitation for a three-phase 12/8 SRM; the
studies also reveal that the maximum torque can be increased.
The flux reversal has a significant impact on the core loss of the SRM. Research on
reducing flux reversal, such as [12–16], was implemented by modifying the shape of the
stator. In this paper, the method to reduce the flux reversal will be discussed based on the
winding arrangement.
The objective of this paper is to investigate the winding arrangements of a three-phase
12/8 SRM under unipolar operation by considering the torque density comparison based
on the same conduction loss, including the core loss effects, the total winding weights, and
torque ripple of different winding arrangements. The simple method, which is the flux
reversal pattern, for evaluating the core loss of the stator in each winding arrangement is
introduced. Furthermore, the winding selection procedure for a three-phase 12/8 SRM
to achieve high torque density, low torque ripple, and cost-saving for manufacturing is
described. The finite element program (JMAG) and PC–SRD software are used mainly
for the calculation of the torque density (static and dynamic), core loss, and structure of
winding; winding weight, slot–fill factor, etc.

2. Winding Arrangement
2.1. Short-Pitched Winding
Figure 1 shows three diagrams of the short-pitched winding configuration and flux paths
of the three-phase 12/8 SRM when a single phase (phase A) is excited with unipolar operation.
For the short-pitched winding, each coil is wound around a single stator pole. The wind-
ing’s polarities are arranged in the stator to give a magnetic pole when each phase is excited,
resulting in the stator poles acting as a north (N) or south (S) magnetic pole, which is the
main cause of magnetic–flux path direction. The 12/8 SRM’s winding polarities of short-
pitched winding are modified to change the magnetic poles. In a particular phase, Figure 1a
shows the magnetic–flux paths when the sequential magnetic poles are the same, e.g., N–
N–N–N and S–S–S–S. The sequential magnetic poles are such that two poles have the same
polarity, e.g., S–N–N–S and N–S–S–N in Figure 1b. The opposite of sequential magnetic
poles, e.g., N–S–N–S and S–N–S–N, is shown in Figure 1c.

$ $ $ $ $ $
$

$
$

$
$

$

$

$
$

$
$

$

$ $ $ $ $ $

D E F

Figure 1. Winding arrangement for single-phase excitation; (a) Short-pitched winding 1; (b) Short-
pitched winding 2; and (c) Short-pitched winding 3.

96
Energies 2022, 15, 284

The torque of SRM with short-pitched winding is produced due to the self-inductance
variation. The mutual-inductance between the phase windings is ineffective and therefore
neglected, resulting in

1 2 dL a 1 dL 1 dLc
T= i + i2b b + i2c (1)
2 a dθ 2 dθ 2 dθ
where the subscripts a, b, and c denote the phase; i, L, and θ are the phase current, self-
inductance, and rotor position, respectively.

2.2. Fully-Pitched Winding


Three diagrams of the winding configuration and flux paths in a three-phase 12/8
SRM when two phases (phase A and B) are excited with unipolar operation are shown in
Figure 2. In the fully-pitched winding arrangement of the three-phase 12/8 SRM, each
coil spans the surrounding three adjacent stator poles. Figure 2a shows the fully-pitched
winding diagram when the number of turns per pole is the same as the number of turns
per pole of the short-pitched winding, while Figure 2b shows the fully-pitched winding
when it is wound with double the number of turns per pole of the short-pitched winding.
The fully-pitched winding can operate with either unipolar (double-phase on at a time) or
bipolar operation (double-phase on at a time and all three phases on at a time) [5–7].

$ $ %
%

$
%

%

$

%

$ % $

D E
Figure 2. Winding arrangement for double-phase excitation; (a) fully-pitched winding 1 and (b) fully-
pitched winding 2.

The torque of the fully-pitched winding is only produced due to the rate of change
of the mutual-inductance (M) among phases, and torque production can be achieved by
exciting two phases simultaneously instead of only one phase in the short-pitched winding.
The self-inductance of an individual phase is constant and has the same value as the
maximum phase inductance at any rotor position. Therefore, the self-inductance of the
fully-pitched winding is independent of the rotor position and is ineffective for torque
production. The torque equation of the fully-pitched winding can be expressed by

dMab dMbc dMca


T = i a ib + ib ic + ic i a (2)
dθ dθ dθ

3. Characteristics of the Winding Arrangement


3.1. Winding Weight and Comparison Based on Equal Conduction Loss
Table 1 shows the dimensions of the three-phase 12/8 SRM. Table 2 shows the winding
details of each winding configuration. The number of turns per pole of the short-pitched
winding and fully-pitched winding 1 is the same, but it doubles in the fully-pitched winding
2. The number of windings of the short-pitched winding for the three-phase 12/8 SRM

97
Energies 2022, 15, 284

generally is the same as the number of stator poles, at 12 coils, as it is a concentrated


winding, while the number of windings in fully pitched is decreased to six coils as it
is a distributed winding. In terms of the slot-fill factor, which is the ratio of the actual
cross–section of copper in a stator slot to the total stator slot area, the fully-pitched winding
1 has the lowest slot–fill factor, while other winding arrangements have the same slot-fill
factor.

Table 1. Details and dimensions of the 12/8 SRM.

Parameter Value
Stator outer diameter [mm] 190
Stator inner diameter [mm] 168
Stator pole arc [mechanical degree] 15
Rotor pole arc [mechanical degree] 16
Shaft diameter [mm] 35
Air gap [mm] 0.3
Stack length [mm] 120
Wire diameter [mm] 0.914 (SWG 20)
Number of strands 2
Rated speed [r/min] 500
Rated torque [Nm] 19
DC link voltage [V] 310
Converter topology Asymmetric bridge

Table 2. Winding arrangement details.

Winding Arrangement Number of Turns per Pole Number of Coils Excitation Type
Short-pitched 1 97 12 Single phase
Short-pitched 2 97 12 Single phase
Short-pitched 3 97 12 Single phase
Fully-pitched 1 97 6 Double-phase
Fully-pitched 2 194 6 Double-phase

The total winding weight of the fully-pitched winding 2 is the highest at 9.0 kg, while
the total weight of windings in the fully-pitched winding 1 is the lowest at 4.5 kg. The slot-
fill factor, total weight of windings, and phase resistance of each winding arrangement are
shown in Table 3. The total winding weight of all short-pitched winding arrangements is
the same at 5.1 kg. The fully-pitched winding 2 has the highest phase resistance value.

Table 3. Electrical and mechanical details.

Winding Arrangement Slot-Fill Factor Total Weight of Windings [kg] Phase Resistance [Ω]
Short-pitched 1 0.37 5.1 1.9
Short-pitched 2 0.37 5.1 1.9
Short-pitched 3 0.37 5.1 1.9
Fully-pitched 1 0.18 4.5 1.7
Fully-pitched 2 0.37 9.0 3.4

In the static machine test, it was thought fair to compare the various designs on the
assumption of equal overall winding loss in each situation. Obviously, iron loss will be a
key issue in the rotating machine: the influence of magnetic flux reversals on iron loss is
explored in a subsequent section, but for a machine rotating at low speed, a comparison
based on equal winding loss is sufficient. The winding loss per phase equation in the
short-pitched winding SRM is given in

1 2
Loss SP = I R SP (3)
3 SP

98
Energies 2022, 15, 284

It is assumed that each phase conducts approximately one-third of an electrical cycle.


Similarly, in the fully-pitched winding, it can be assumed that each phase conducts approx-
imately two-thirds of one electrical cycle (two-phase excitation), and then the winding loss
per phase equation in the fully-pitched winding SRM is given in (4):

2 2
Loss FULL = I R FULL (4)
3 FULL
where the subscripts SP and FULL denote the short-pitched winding and fully-pitched
winding, respectively; I and R are the conduction current and phase resistance, respectively.
To compare the five winding arrangements of a three-phase 12/8 SRM based on
the same conduction loss for the same motor structure, the phase current of each wind-
ing arrangement is selected according to Equations (3) and (4) for the short-pitched and
fully-pitched winding, respectively. Table 4 shows the phase current of each winding
configuration based on equal conduction loss. In the Finite Element Analysis (FEA) model,
the phase current of each winding arrangement in Table 4 is employed to simulate phase
torque production.

Table 4. Phase winding current comparison for the same conduction loss.

Winding Arrangement Phase Winding Current at the Same Conduction Loss [A]
Short-pitched 1, 2, 3 15
Fully-pitched 1 11.27
Fully-pitched 2 7.98

The excitation methods under unipolar operation of the winding arrangement are
shown in Table 2. The short-pitched winding arrangements are excited with single-phase
excitation at a time, while the fully-pitched winding arrangements are excited with double-
phase excitation at a time.

3.2. Core Loss Based on Flux Reversals


The core losses consist of three parts: hysteresis loss, eddy current loss, and excess
loss as shown in Equation (5). The calculation of the core loss per unit volume (Pv ) in the
frequency domain is widely used for ferromagnetic material [17] as it can be expressed
as Equation (6). k hys , k eddy , k excess are the coefficients and β is the parameter of the core
loss which are obtained from manufacturer data or experimental testing based on the
curve–fitting method. Bm is the peak magnetic flux density and f is the frequency. This
equation is suitable for sinusoidal flux density waveforms, and it does not consider the
minor loop in the main hysteresis loop.The frequency of the flux is determined by the phase
switching frequency:

Pv = Phys + Peddy + Pexcess (5)


β
Pv = k hys f Bm + k eddy ( f Bm )2 + k excess ( f Bm )1.5 (6)
For SRM, the flux, which is distributed on the machine, is nonsinusoidal waveforms.
Using the conventional equation may lead to large errors in the core loss calculation. Then,
the core loss can be analyzed in the time domain [18], which is suitable for the nonsinusoidal
flux waveform, and it can be given by Equations (7)–(9):

dB
Phys (t) = k hys Hirr (7)
dt
 2
1 dB
Peddy (t) = k eddy 2 (8)
2π dt

99
Energies 2022, 15, 284

 1.5
1  dB 
Pexcess (t) = k excess   (9)
Ce dt
Hirr is evaluated based on an equivalent elliptical loop and Ce is the numerical inte-
gration parameter as described in [18]. From the core loss equation in the time domain, it
can be seen that the main contributory impact of the core loss is the rate of change of the
flux density with time; dB/dt. Modifying winding arrangement can change the direction
of the flux path, which affects the flux density of the machine. The phenomenon of flux
reversals in the stator core back is one of the causes that affects the variation of the flux
density in the SRM [14–16]. Therefore, minimizing the number of flux reversals occurring
on the stator core back can decrease the core loss in SRM.

4. Selection of Winding Arrangement


4.1. Average Static Phase Torque
The effective static phase torque is generated according to the inductance profile of
each winding arrangement, as shown in Figure 3. The effective torque regions, which are the
positive torque production zones of the three-phase 12/8 SRM, are 22.5 mechanical degrees.
Figure 4 shows the effective static phase torque profiles of each winding arrangement
based on equal conduction loss. From Figure 4, the fully-pitched winding 2, which is
double-phase excitation, can produce the highest peak torque, at 44.08 Nm. For short-
pitched winding arrangements, the short-pitched winding 1, which requires single phase
excitation, can generate a peak torque of 35.18 Nm, but it produces negative torque at some
position, and the short-pitched windings 2 and 3, which require single phase excitation,
can generate a peak torque of 40.47 Nm and 41.4 Nm, respectively. In other fully-pitched
winding arrangements, the fully-pitched winding 1, which uses double-phase excitation,
can produce a peak torque of 29.3 Nm.
It can be observed in the short-pitched winding 1, which produces some negative
torque due to its inductance profile having an early decrease slope before going to an
aligned position at 22.5 degrees. The inductance profile of the fully-pitched winding 2
has a high slop caused by the high number of winding turns at 194 compared with other
winding arrangements. Table 5 lists the average value of the effective torque on the basis of
the same conduction loss according to the torque profiles in Figure 4. It is clear that the
fully-pitched winding 2 can generate the highest average torque, while the short-pitched
winding 1 can produce the lowest average torque.

0.25 0.008 0.5


0.007 0.45
0.2 0.4
0.006
Inductance [H]

Inductance [H]

Inductance [H]

0.35
0.15 0.005 0.3
SP 1 0.004 0.25
0.1 0.2 FP 1
SP 2 0.003 SP 2
0.15 FP 2
SP 3 0.002
0.05 0.1
0.001 0.05
0 0 0
0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90
Rotor Position [deg] Rotor Position [deg] Rotor Position [deg]

D E F

Figure 3. Inductance profiles of a three-phase 12/8 SRM at different winding arrangements; (a) Short-
pitched winding; (b) Short-pitched winding 2; and (c) Fully-pitched winding.

100
Energies 2022, 15, 284



6KRUWSLWFK
 6KRUWSLWFK
6KRUWSLWFK
)XOOSLWFK
 )XOOSLWFK

6WDWLFWRUTXH>1P@






 
      
5RWRUSRVLWLRQ>GHJ@

Figure 4. Effective static torque of a three-phase 12/8 SRM at different winding arrangements on the
basis of equal conduction loss.

Table 5. Average static torque at the same conduction loss.

Winding Arrangement Static Torque [Nm] Torque/Copper Weight [Nm/kg]


Short-pitched 1 9.8 1.9
Short-pitched 2 19.6 3.8
Short-pitched 3 21.6 4.2
Fully-pitched 1 15.9 3.5
Fully-pitched 2 23.4 2.6

4.2. Number of Flux Reversals on Stator Yoke


In this paper, the number of flux reversals is considered only on the stator yoke.
The leakage flux is neglected. The yoke of the 12/8 SRM stator is divided into 12 segments.
Each segment of the stator yoke experiences flux passing through it when the phase winding
is excited. With different phases of winding excitation, the direction of the flux path in each
stator yoke segment is varied.
As mentioned in the above discussion, the rate of change of the flux density with time
significantly affects the core loss of the SRM stator. The number of flux reversals is one of
the causes that involves the rate of change of the flux density on the stator core. In this
paper, the number of flux reversals occurring on the stator yoke is observed by counting the
number of changes in flux path direction in each stator segment when the phase windings
are sequentially excited in one revolution. The number of flux reversals in the stator yoke
can be given as
 n 
NFR = ∑ NFD (i ) ∗ Ns (10)
i =1

where NFR is the total number of flux reversals on the yoke of the stator, NFD is the
number of changing flux path directions in each segment of the stator yoke when the phase
windings are sequentially excited (A–B–C–A) , Ns is the number of stator poles, and n is
the number of stator yoke segments.
Table 6 shows the number of flux reversals on the stator yoke in the different winding
configurations. The short-pitched winding 3 and the fully-pitched winding have the highest
number of flux reversals on the stator yoke at 288, while the short-pitched winding 2 does
not have the number of flux reversals on the stator yoke. The short-pitched winding 1 still
has a high number of flux reversals on the yoke of the stator at 240.

101
Energies 2022, 15, 284

Table 6. Number of flux reversals in the stator yoke in 1 revolution.

Winding Arrangement Number of Flux Reversals


Short-pitched 1 240
Short-pitched 2 0
Short-pitched 3 288
Fully-pitched 1 288
Fully-pitched 2 288

4.3. Consideration of Winding Selection


In terms of application, the ventilation exhaust fan for poultry farming requires a
motor to run at a low rotational speed, have a high starting torque output, and be robust
enough for the hard environment in a farming system. In terms of motor manufacturing,
simple structure, no complexity in the assembly process, and low copper consumption for
motor winding are the key factors for the candidate motor. From both points of view, SRM
is a suitable choice for this application. Rated speed and load torque, including voltage
range and converter topology of the SRM in this research, are shown in Table 1.
The winding selection of SRM is determined by the characteristics of creating the high
torque capacity to achieve a high–efficiency motor and using low copper, resulting in cost
savings in the manufacture. Tables 2 and 3 are used for considering the complexity of
the winding assembly process and copper consumption in each winding configuration.
The fully-pitched winding has a small number of coils at 6 coils, while the short-pitched
winding has 12 coils. However, in terms of the manufacturing process, the short-pitched
winding with a concentrated winding gives more advantages, such as the winding of a coil
around a single–pole and short end–turn of coil compared with the fully-pitched (winding
of a coil around three stator poles), which is similar to a distributed winding. Furthermore,
the fully-pitched winding 2 uses more copper at 9 kg compared with any other type of
winding arrangement.
To compare the torque performance, the ratio of the static phase torque per copper
weight is applied for consideration as the stator and rotor core are the same weight. The
static phase torque results based on equal conduction loss were obtained from the FEA
model of SRM, which applied the phase current value in Table 4 calculating based on equal
conduction loss according to Equations (3) and (4) for the short-pitched and fully-pitched
winding, respectively.
As shown in Table 5, the ratio of the static phase torque per copper weight of the
short-pitched windings 2 and 3 is high compared with other types of winding configuration.
In terms of motor efficiency, the number of flux reversals of the short-pitched winding 2
is the lowest. On the other hand, the short-pitched winding 3 has a high number of flux
reversals, as shown in Table 6. Therefore, two candidates for the winding arrangement of
the 12/8 SRM are the short-pitched windings 2 and 3. In the next analysis, the dynamic
simulation analysis of each winding pattern will be discussed.

5. Dynamic Simulation
The dynamic simulation in this section is performed by JMAG finite element analysis
software. The control parameters of the 12/8 SRM in the dynamic simulation are the
same. The advance turn–on angle is 0 degrees, the conduction period is 22.5 degrees in the
short-pitched winding, the conduction period is 30 in the fully-pitched winding, and the
rotational speed is 500 r/min. A conventional hysteresis current control is applied and the
maximum current is 9 A. The five winding configurations of the 12/8 SRM are simulated.
Average torque and torque ripple are the essential considerations in this simulation.
Figure 5 shows the flux density and flux path of the dynamic simulation. The result of
the dynamic simulation is shown in Figures 6 and 7 and comparative results in Table 7. The
dynamic torque performance of the double-phase excitation of the fully-pitched winding
is shown in Figure 7. The torque ripple rate of the fully-pitched winding 1 and the fully-
pitched winding 2 is 1.3 and 1.0, respectively. The average torque of the fully-pitched

102
Energies 2022, 15, 284

winding is the highest, at 36.55 Nm. However, with double-phase excitation and high
phase resistance of the fully-pitched winding, it produces more copper losses compared
with the other winding arrangements.

Table 7. Comparison of dynamic simulation performance.

Winding Arrangement Average Torque [Nm] Torque Ripple Rate


Short-pitched 1 3.54 10.28
Short-pitched 2 11.81 2.25
Short-pitched 3 19.0 0.66
Fully-pitched 1 17.12 1.30
Fully-pitched 2 36.55 1.00

Figure 5. Leakage-flux path of the short-pitched winding 2 when operates at high current densities.

D E F

Figure 6. Torque and current waveform (Phase A) for single-phase excitation; (a) Short-pitched
winding 1; (b) Short-pitched winding 2; and (c) Short-pitched winding 3.

103
Energies 2022, 15, 284

60 60
50 50
40 40

Torque [Nm]

Torque [Nm]
30 30
20 20
10 10
0 0
-10 0 0.01 0.02 0.03 0.04 0.05 0.06 -10 0 0.01 0.02 0.03 0.04 0.05 0.06
-20 -20
-30 -30
Time [s] Time [s]

20 20
Phase A Phase A
15 Phase B 15

Current [A]
Current [A]

Phase B

10 10

5 5

0 0
0 0.01 0.02 0.03 0.04 0.05 0.06 0 0.01 0.02 0.03 0.04 0.05 0.06
Time [s] Time [s]

D E

Figure 7. Torque and current waveform for double-phase excitation; (a) Fully-pitched winding 1 and
(b) Fully-pitched winding 2.

The dynamic torque performance of the single-phase excitation in Figure 6 shows that
the short-pitched winding 1 has a dramatically torque ripple rate of 10.28, and the average
torque is lowest at 3.54 Nm. The short-pitched winding 2 has a torque ripple rate of 2.25
and the average torque decreases to 11.81 Nm. compared with the short-pitched winding
3 having a low torque ripple rate at 0.66 and average torque at 19 Nm. To describe this
phenomenon, Figure 5 shows the flux density and flux path of the short-pitched winding
2. The simulation results of the short-pitched winding 2 show that the leakage–flux paths
appear on the adjacent stator poles when operating at high torque density. Therefore,
the number of flux reversals in the stator yoke increases and generates more negative
torque during phase commutation. The leakage–flux paths are shown in the red dot circle
in Figure 5. The short-pitched winding 2 has good torque performance and reduces the
number of flux reversals when operating at low torque density, as was revealed in [15].
For this reason, the average torque of the short-pitched winding 2 decreases, and the
performance of the SRM drops to 76.63%, while the short-pitched winding 3 has a higher
average torque and efficiency at 84.9%. However, the core loss of the short-pitched winding
2 is still lower than the short-pitched winding 3. Table 8 reveals comparative results of
dynamic simulation performance between short-pitched 2 and short-pitched 3 in terms
of core loss at the stator and efficiency performance. For ventilation fan applications, the
torque ripple has a significant impact on this application. A high torque ripple will lead to
increased vibration and acoustic noise.

Table 8. Comparison of motor performance between short-pitched 2 and short-pitched 3.

Short-Pitched 2 Short-Pitched 3
Average torque [Nm] 11.81 19.0
Torque ripple rate 2.25 0.66
Core loss at stator [W] 37.67 38.44
Efficiency [%] 76.63 84.9

From the dynamic simulation results, it can be concluded that the short-pitched
winding 3 is suitable for 12/8 SRM in terms of cost-saving, simple manufacturing process,
torque ripple, and motor performance.

104
Energies 2022, 15, 284

6. Flux Reversal Reduction


Although the short-pitched winding 3 gives a better performance for 12/8 SRM, the
number of flux reversals of this winding configuration is still high. This paper proposes an
analysis method that can reduce the flux reversals on the stator yoke. The short-pitched
winding 3 is divided into Design A and Design B as follows:
• In Design A, as shown in Figure 8, the sequential magnetic poles in a clockwise
direction are SNSN when excited phase A, NSNS when excited phase B, and NSNS
when excited phase C.
• In Design B, as shown in Figure 9, the sequential magnetic poles in a clockwise
direction are SNSN when excited phase A, NSNS when excited phase B, and SNSN
when excited phase C.

     

     
$ $ %
&
%
&

&
%

     

&
%
$

$
$

$

%

&
     
%

&
&
%
&
$ $ %
     

     
D E F
Figure 8. Excitation pattern and flux-path of the short-pitched winding 3 in Design A; (a) Phase A is
excited; (b) Phase B is excited; and (c) Phase C is excited.

     

     
$ $ %
&
%
&
&
%

     
&
%
$

$
$

$

%

&

     
%

&

&
%
&
$ $ %
     

     
D E F
Figure 9. Excitation pattern and flux-path of the short-pitched winding 3 in Design B; (a) Phase A is
excited; (b) Phase B is excited; and (c) Phase C is excited.

The number of flux reversals in each segment of the stator yoke in the clockwise (A–B–
C–A) and counter-clockwise (A–C–B–A) direction of Design A is illustrated in Table 9. To
calculate the total number of flux reversals in one revolution, Equation (10) is used. The total
number of flux reversals on the yoke of the stator is 192 in both directions (clockwise and
counter–clockwise). Using the same procedure, Design B has a total number of flux reversals
in one revolution of 288 in both directions (clockwise and counter–clockwise) as shown in
Table 10. Table 11 reveals the results of the comparison of the short-pitched winding 3 with

105
Energies 2022, 15, 284

Designs A and B. The results show that the stator core loss is decreased in the short-pitched
winding 3 with Design A.

Table 9. Number of flux reversals in Design A.

No. of
Stator Flux-Linkage No. of Flux-Linkage Reverse
Yoke Direction Reverse Direction Direction
Sector (Clockwise) Direction (Counter (Counter
(Clockwise) Clockwise) Clockwise)
1 →←←→ 2 →←←→ 2
2 ←←←← 0 ←←←← 0
3 ←←→← 2 ←→←← 2
4 ←→→← 2 ←→→← 2
5 →→→→ 0 →→→→ 0
6 →→←→ 2 →←→→ 2
7 →←←→ 2 →←←→ 2
8 ←←←← 0 ←←←← 0
9 ←←→← 2 ←→←← 2
10 ←→→← 2 ←→→← 2
11 →→→→ 0 →→→→ 0
12 →→←→ 2 →←→→ 2
Total 16 Total 16

Table 10. Number of flux reversals in Design B.

No. of
Stator Flux-Linkage No. of Flux-Linkage Reverse
Yoke Direction Reverse Direction Direction
Sector (Clockwise) Direction (Counter (Counter
(Clockwise) Clockwise) Clockwise)
1 →→←→ 2 →←→→ 2
2 ←→←← 2 ←←→← 2
3 ←→→← 2 ←→→← 2
4 ←←→← 2 ←→←← 2
5 →←→→ 2 →→←→ 2
6 →←←→ 2 →←←→ 2
7 →→←→ 2 →←→→ 2
8 ←→←← 2 ←←→← 2
9 ←→→← 2 ←→→← 2
10 ←←→← 2 ←→←← 2
11 →←→→ 2 →→←→ 2
12 →←←→ 2 →←←→ 2
Total 24 Total 24

Table 11. Comparison of winding performance short-pitched winding 3 Design A (SP 3 Design A)
and short-pitched winding 3 Design B (SP 3 Design B).

SP 3 Design A SP 3 Design B
Average torque [Nm] 18.9 19.0
Torque ripple rate 0.66 0.66
Core loss at stator [W] 36.36 38.44
Efficiency [%] 85.14 84.9

106
Energies 2022, 15, 284

Figure 10 shows the comparative results of core loss between short-pitched 3 Design A
and Design B at each rotational speed and current level, respectively. It can be seen that Design
A can reduce the core loss by around 7–9% at the rotational speed of 300–400 r/min with low
current and around 10–13% at the rotational speed of 500 r/min with high current.

100 100
90 90
80 80
Core loss [W]

Core loss [W]


70 70
60 60
50 50
40 40
30 30
20 Design A 20 Design A
10 Design B 10 Design B
0 0
200 300 400 500 600 0 5 10 15 20
Rotational speed [r/min] Current [A]
D E

Figure 10. Core loss of the short-pitched winding 3 Design A and Design B; (a) at different rotational
speeds and (b) at different current levels.

7. Experimental Results
The three-phase 12/8 SRM prototype was built for ventilation fan applications using
the existing induction motor frame size of 112 M, including cover, motor frame, and shaft
according to the IEC standard. Therefore, the maximum outer diameter is 190 mm and
the shaft diameter is 35 mm. Figure 11 reveals the example of a single coil winding on the
stator pole. The winding arrangement of the SRM prototype is the short-pitched winding
3 Design A, and the winding diagram is shown in Figure 12. Figure 13 illustrates the
completed prototype of the 12/8 SRM.

Figure 11. One coil winding on the stator pole.

3KDVH$  3KDVH %  3KDVH&

Figure 12. Winding diagram of the 12/8 SRM prototype.

107
Energies 2022, 15, 284

Figure 13. Stator, rotor, and completed 12/8 SRM prototype.

The experimental system setup is shown in Figure 14. The rotor of the three-phase 12/8
SRM was connected to the rotor of the load machine (20 kW Siemens induction motor) via a
flexible coupling and torque transducer. The turn-on and turn-off angles of SRM are 0 and
22.5 deg. The DC link voltage was 310 V. The experimental test was implemented at 500 r/min.
The rated load torque output of 19.4 Nm of the load machine was applied to the 12/8 SRM.
The current and voltage waveform of the experimental test are shown in Figure 15a,b at the
rotational speed of 600 r/min with load torque at 18.5 Nm of the load machine. It can be seen
that the SRM prototype running at 600 r/min can carry less load torque due to the influence
of voltage back electromotive force, which increases with rotational speed. Therefore, phase
current cannot reach the level of the desired current, leading to low torque capability. However,
when the phase current decreases, the loss of the copper also reduces. For this reason, the
motor performance of 600 r/min gives high efficiency.

7RUTXH
650 WUDQVGXFHU /RDGPDFKLQH

Figure 14. Experimental test rig.

Figure 16 shows the experimental results of torque vs. speed and efficiency vs. speed
curves with different rotational speeds between 300–1500 r/min. The maximum torque
of the SRM prototype is 21.8 Nm at 300 r/min. At the rated speed of 500 r/min, the SRM
prototype can carry torque at 19.4 of load torque which covers the desired rated torque,
and the efficiency of the motor is 82.5%. The maximum efficiency of the SRM prototype
can reach 83.1% at 600 r/min with 18.5 Nm of load torque.

108
Energies 2022, 15, 284

a b

Figure 15. Phase A current (Top) and voltage (Bottom) waveform of the 12/8 SRM prototype using
SP 3 Design A winding configuration; (a) at 500 r/min with 19.4 Nm of load torque and (b) at
600 r/min with 18.5 Nm of load torque.

The comparative results at rated conditions between the simulation model and experi-
mental test are illustrated in Table 12. The SRM prototype gives a torque output of 19.4 Nm
close to the simulation model, which has 18.9 Nm. However, the efficiency of the SRM proto-
type decreased to 82.5%, while the simulation model could achieve 85.14%. In the experimental
test, the maximum current of the hysteresis current control is higher than that of the simulation,
so this will lead to an increase in the copper loss of the system. To achieve the rated torque of
19.0 Nm, the SRM prototype requires more phase current than that in the simulation model
due to the imperfections during the manufacture and assembly process, such as errors in air
gap length, non-uniform air gap, and eccentricity error, etc.

100 25 21.8
20.5
95 19.4
18.5
20
90 16.1
Torque [Nm]
Efficiency [%]

82.5 83.1 82.4 82.8 82.4 81.7


85 80.2 79.4 15
11.3
80 77.1 76.6
75.1 74.2
73.3 10 8
75 5.9
70 4.4
5 3.3 2.7
2.2 2
65
60 0
300 500 700 900 1100 1300 1500 300 500 700 900 1100 1300 1500
Rotational speed [r/min] Rotational speed [r/min]

Figure 16. Efficiency and torque at different rotational speeds.

Table 12. Experimental results.

SP 3 Design A FEA SP 3 Design A Exp


Average torque [Nm] 18.9 19.4
Efficiency [%] 85.14 82.5
Aligned inductance [mH] 218.57 215.54
Unaligned inductance [mH] 21.86 27.37

8. Conclusions
This research has introduced the criteria to consider for choosing the winding configu-
ration of three-phase 12/8 SRM as follows:
In the analysis of the static characteristic: (1) the ratio value of static phase torque
(based on equal conduction loss) per weight of copper winding needs to be considered
for selecting winding configuration to produce high efficiency and cost-saving in the
production process; (2) the number of flux reversals on the stator yoke of the SRM, winding
configuration having a low number of flux reversals can reduce the core loss.

109
Energies 2022, 15, 284

In the analysis of the dynamic characteristics, the torque ripple is the main parameter
that each winding configuration needs to consider, especially since the torque ripple has a
significant impact on the ventilation fan application.
The comparison of the winding arrangements for the same motor structure reveals
that the total winding weight of the fully-pitched winding 2 is the highest because of the
increased end windings. The comparison of the average torque results on the basis of equal
conduction loss shows that the fully-pitched winding 2 with double-phase unipolar excita-
tion and the short-pitched winding 3 with single phase unipolar excitation are effective for
enhancing the torque density. The current level of the fully-pitched winding 2 has to be
reduced to attain the same conduction loss where average output torque is still higher than
any other winding arrangement.
The consideration of winding selection is discussed based on cost-saving, the com-
plexity of the manufacturing process, and machine performance, including torque output
and torque ripple. The short-pitched winding 3 gives significantly better performance
in terms of the ratio of the static phase torque per copper weight compared with other
winding configurations. For initial consideration of core loss in SRM, the counting of the
number of flux reversals on the stator yoke is proposed. The simulation results confirm
that minimizing the number of flux reversals in the stator yoke by modifying the winding
polarities can decrease the core loss of the SRM. Finally, the 12/8 SRM prototype was built
and tested (load test and characteristic test).

Author Contributions: Conceptualization, R.P. and K.T.; methodology, R.P. and K.T.; validation, R.P.,
S.K., P.J., P.S. and K.T.; formal analysis, R.P. and K.T.; investigation, R.P., P.S. and K.T.; resources, S.K.,
P.J. and P.S.; writing—original draft preparation, R.P.; writing—review and editing, R.P., S.K., P.J.,
P.S. and K.T.; visualization, R.P. and S.K.; supervision, K.T. All authors have read and agreed to the
published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

Nomeclatures

N north magnetic pole


S south magnetic pole
T torque
i phase current
θ rotor position
L self-inductance
M mutual-inductance
SP short-pitched winding
FULL fully-pitched winding
I conduction current
R phase resistance
Pv core losses
Phys hysteresis loss
Peddy eddy current loss
Pexcess excess loss
k hys hysteresis loss coefficient
k eddy eddy current loss coefficient
k excess excess loss coefficient

110
Energies 2022, 15, 284

β core loss parameter based on the curve-fitting method


Bm peak magnetic flux density
B magnetic flux density
Hirr evaluated value based on an equivalent elliptical loop [18]
Ce numerical integration parameter [18]
NFR total number of flux reversals on stator yoke
NFD number of changing flux path direction in each segment of stator yoke
Ns number of stator poles
n number of stator yoke segments

References
1. Miller, T.J.E. Switched Reluctance Motors and Their Control; Magna Physics Publishing and Clarendon Press: Oxford, UK, 1993.
2. Krishnan, R. Switched Reluctance Motor Drives: Modeling, Simulation, Analysis, Design, and Applications; CRC Press: Boca Raton, FL,
USA, 2001.
3. Widmer, J.D.; Martin, R.; Mecrow, B.C. Optimization of an 80-kW Segmental Rotor Switched Reluctance Machine for Automotive
Traction. IEEE Trans. Ind. Appl. 2015, 51, 2990–2999. [CrossRef]
4. Rahman, M.S.; Lukman, G.F.; Hieu, P.T.; Jeong, K.I.; Ahn, J.W. Optimization and Characteristics Analysis of High Torque Density
12/8 Switched Reluctance Motor Using Metaheuristic Gray Wolf Optimization Algorithm. Energies 2021, 14, 2013. [CrossRef]
5. Mecrow, B. Fully pitched-winding switched-reluctance and stepping-motor arrangements. IEE Proc. B (Electr. Power Appl.) 1993,
140, 61–70. [CrossRef]
6. Barrass, P.G.; Mecrow, B.C.; Clothier, A.C. Bipolar operation of fully-pitched winding switched reluctance drives. In Proceedings
of the 1995 Seventh International Conference on Electrical Machines and Drives (Conf. Publ. No. 412), Durham, UK, 11–13
September 1995; pp. 252–256. [CrossRef]
7. Mecrow, B.C. New winding configurations for doubly salient reluctance machines. IEEE Trans. Ind. Appl. 1996, 32, 1348–1356.
[CrossRef]
8. Li, Y.; Tang, Y. Switched reluctance motor drives with fractionally-pitched winding design. In Proceedings of the PESC97, Record
28th Annual IEEE Power Electronics Specialists Conference, Formerly Power Conditioning Specialists Conference 1970–71. Power
Processing and Electronic Specialists Conference 1972, St. Louis, MO, USA, 27 June 1997; Volume 2, pp. 875–880. [CrossRef]
9. Tang, Y. Switched reluctance motor with fractionally pitched windings and bipolar currents. In Proceedings of the Conference
Record of 1998 IEEE Industry Applications Conference, Thirty-Third IAS Annual Meeting (Cat. No. 98CH36242), St. Louis, MO,
USA, 12–15 October 1998; Volume 1, pp. 351–358. [CrossRef]
10. Chen, H.; Song, Q. Windings arrangement of a three-phase switched reluctance machine. In Proceedings of the IEEE International
Electric Machines and Drives Conference, IEMDC’03, Madison, WI, USA, 1–4 June 2003; Volume 3, pp. 1665–1668. [CrossRef]
11. Pupadubsin, R.; Chayopitak, N.; Karukanan, S.; Champa, P.; Somsiri, P.; Tungpimolrut, K. Comparison of winding arrangements
of three phase switched reluctance motor under unipolar operation. In Proceedings of the 2012 15th International Conference on
Electrical Machines and Systems (ICEMS), Sapporo, Japan, 21–24 October 2012; pp. 1–4.
12. Augustine, M.; Balaji, M.; Kamaraj, V. Characteristics Assessment of Switched Reluctance Motor with Segmented Rotor. In
Proceedings of the 2019 IEEE 1st International Conference on Energy, Systems and Information Processing (ICESIP), Chennai,
India, 4–6 July 2019; pp. 1–6. [CrossRef]
13. Prabhu, S.; Balaji, M.; Kamaraj, V. Analysis of two phase switched reluctance motor with flux reversal free stator. In Proceedings
of the 2015 IEEE 11th International Conference on Power Electronics and Drive Systems, Sydney, NSW, Australia, 9–12 June 2015;
pp. 320–325. [CrossRef]
14. Oh, S.; Krishnan, R. Two-Phase SRM With Flux-Reversal-Free Stator: Concept, Analysis, Design, and Experimental Verification.
IEEE Trans. Ind. Appl. 2007, 43, 1247–1257. [CrossRef]
15. Lobo, N.S.; Swint, E.; Krishnan, R. M-Phase N-Segment Flux-Reversal-Free Stator Switched Reluctance Machines. In Proceedings
of the 2008 IEEE Industry Applications Society Annual Meeting, Edmonton, AB, Canada, 5–9 October 2008; pp. 1–8. [CrossRef]
16. Seshadri, A.; Lenin, N.C. Review based on losses, torque ripple, vibration and noise in switched reluctance motor. IET Electr.
Power Appl. 2020, 14, 1311–1326. [CrossRef]
17. Bertotti, G. General properties of power losses in soft ferromagnetic materials. IEEE Trans. Magn. 1988, 24, 621–630. [CrossRef]
18. Lin, D.; Zhou, P.; Fu, W.N.; Badics, Z.; Cendes, Z.J. A dynamic core loss model for soft ferromagnetic and power ferrite materials
in transient finite element analysis. IEEE Trans. Magn. 2004, 40, 1318–1321. [CrossRef]

111
energies
Article
Performance Improvement of a Switched Reluctance Motor and
Drive System Designed for an Electric Motorcycle
Seubsuang Kachapornkul, Ruchao Pupadubsin, Pakasit Somsiri, Prapon Jitkreeyarn and
Kanokvate Tungpimolrut *

Machines and Power Conversions Research Team, National Electronics and Computer Technology Center,
112 Thailand Science Park, Phahonyothin Rd., Khlong Nueng, Khlong Luang 12120, Pathumthani, Thailand;
[email protected] (S.K.); [email protected] (R.P.);
[email protected] (P.S.); [email protected] (P.J.)
* Correspondence: [email protected]

Abstract: In this paper, the implementation of a switched reluctance motor (SRM) and drive system
for the propulsion system of a two-seat electric motorcycle is described. The overall design focuses
on the required vehicle speed, acceleration, driving distance, and overall system cost, as well as
reliability. The performance of the three-phase 6/4 pole (six-stator pole and four-rotor pole) and
four-phase 8/6 pole (eight-stator pole and six-rotor pole) are investigated and compared by static
performance analysis and dynamic performance analysis. Their performance is further investigated
by finite element analysis. The indirect torque controller in a drive system for optimal torque and
efficiency operation is also mentioned. A methodology for rotor position detection and its hardware
implementation are also proposed. The designed 3.5 kW three-phase 6/4 pole SRM and its drive
system were constructed and tested on the test bench. A maximum efficiency of about 82% could be
achieved for the SRM and drive system. It was also installed on a 120-cc electric motorcycle, and
Citation: Kachapornkul, S.;
the vehicle’s performance was also validated by on-road and dynamometer testing. The maximum
Pupadubsin, R.; Somsiri, P.; vehicle speed reached was 82 km/h, and a cruising distance of about 98 km at a constant speed of
Jitkreeyarn, P.; Tungpimolrut, K. 40 km/h was measured.
Performance Improvement of a
Switched Reluctance Motor and Keywords: electric motorcycle; switched reluctance motors; indirect torque control
Drive System Designed for an
Electric Motorcycle. Energies 2022, 15,
694. https://doi.org/10.3390/
en15030694 1. Introduction
Academic Editor: Vladimir Prakht A motorcycle is the most popular vehicle in Thailand and ASEAN countries. In Thailand,
more than 20 million units, or 50% of all on-road vehicles, registered to the Department of
Received: 21 November 2021
Transportation are motorcycles. Furthermore, there has been a rapidly increasing trend in
Accepted: 14 January 2022
the number of motorcycles during the pandemic for delivery services and urban commuting,
Published: 18 January 2022
which has caused more serious air pollution and health problems. In order to tackle
Publisher’s Note: MDPI stays neutral this problem, converting existing old motorcycles or using new electric motorcycles is
with regard to jurisdictional claims in considered to be an interesting policy of government agencies [1]. Recently, some electric
published maps and institutional affil- motorcycles and small electric scooters have been exported. The propulsion systems
iations. of those vehicles mainly use the motors that contain permanent magnets inside, which
causes some difficulties for local mass production and uncontrollable price fluctuation.
From the literature [2,3], the switched reluctance motor (SRM) is one type of motor that
does not contain any rare earth materials or permanent magnets inside. It presents many
Copyright: © 2022 by the authors.
advantages, opportunities, and challenges over the other type of electric motors for electric
Licensee MDPI, Basel, Switzerland.
This article is an open access article
vehicle applications. Various structures and winding configurations of the SRM as well as
distributed under the terms and
various driving topologies have been reported [4]. It has been widely studied and installed
conditions of the Creative Commons for traction drives in various hybrid [5–7] and electric vehicles, including bicycles [8],
Attribution (CC BY) license (https:// forklifts [9], and light electric vehicles [10]. The design, production, and verification of a
creativecommons.org/licenses/by/ complete running prototype based on SRM wheel hub drive train has been reported [11].
4.0/). A control method in the drive systems of electric motorcycles using SRMs [12] has also

Energies 2022, 15, 694. https://doi.org/10.3390/en15030694 113 https://www.mdpi.com/journal/energies


Energies 2022, 15, 694

been previously reported. Some previous publications focused only on the performance
improvement of the SRM drive system by optimizing the turn on and turn off angle at each
rotating speed and vehicle load torque [13], or by simultaneously adjusting the geometry
and commutation angles [14].
The SRM has a special structure and several advantages, such as rugged and low-cost
construction, a simple stator and rotor structure, an easy cooling system, high reliability,
and good performance over a wide speed range. These characteristics make it a low-cost
and rugged alterative for industry applications, a worthy competitor to other drive systems,
and suitable for various applications. Additionally, its performance can be easily adjusted
for various load profiles. Thus, it is also one of the most interesting applications for electric
and hybrid vehicle drive systems. Unlike permanent magnet machines, the SRM is easy to
manufacture with a cheaper cost, since it does not need a permanent magnet. Moreover,
the SRM is very suitable for harsh and high-temperature environments. For the sake of
safety in a traction drive system [2], very high torque in the low-speed region is required.
Furthermore, for the wide constant power region during high vehicle speeds, the power
train must be mostly operated in the constant power region. It is reported in [3] that the
three-phase 6/4 pole SRM has an overload capacity and constant power region with a
longer range, which is suitable for the traction drive application. However, it has a bigger
torque ripple and some vibration. Therefore, some papers overlook this problem by using
a four-phase motor [7] or increasing the number of stator and rotor poles [5], such as 8/6,
12/8, or 6/10 pole [8] SRMs. However, the increasing number of components for increasing
phases also causes the drive circuit to be more expensive and have more complexity.
Until now, there has been a variety of literature on the design of SRMs. The design
in each paper focuses on different points of view. In the case of electric motorcycle appli-
cations, it is very challenging to design a drive system and a high torque density motor
(per volume or weight) with limited space, low cost, and high reliability in order to meet
high performance and high efficiency requirements and to make it acceptable for many
customers. In this paper, the first task to be accomplished is to design a motor to satisfy the
performance specifications and requirements of the vehicle. A high starting torque motor
and drive system was designed in order to reduce the acceleration and deceleration time to
an acceptable value. The motor design mainly focuses on the starting torque, whereas the
other criteria, such as the good performance and high efficiency of the motor for longer
driving distances, are also taken into consideration. The geographical dimensions for the
design guidelines of the stator, rotor core, and winding parameter for the four-phase 8/6
pole and three-phase 6/4 pole SRM were considered. The comparison of both types of
SRMs was investigated by the simulation results using a static performance analysis tool.
The dynamic performance analysis of the four-phase 8/6 pole and three-phase 6/4 SRMs,
including the control methodology in the drive system, were also investigated using the
dynamic simulation tool. Furthermore, the developed torque and flux density of different
winding configurations for the three-phase 6/4 SRM were also investigated. In order to
implement a low-cost and high-reliability drive system, an indirect toque control method
for optimal torque and efficiency operation and higher accuracy of rotor position detection
based on a combination of a magnetic sensing circuit and an initial rotor position estima-
tion method are proposed. The construction and the measured performance on the test
bench of the selected three-phase 6/4 pole SRM and drive system are described. Finally,
the measured results for the vehicle performance on a dynamometer and during on-road
testing are also considered.

2. Design Methodology
2.1. Specification of the Electric Motorcycle
In this paper, the SRM is designed for an electric motorcycle, comparable to a con-
ventional internal combustion engine (ICE) with a motorcycle size of 120 cc. The targeted
specifications of the electric motorcycle for one or two passengers, including maximum
speed, acceleration from 0–30 km/h, and driving range per charge (at a constant speed

114
Energies 2022, 15, 694

of 40 km/h), are indicated in Table 1. This targeted specification is considered one of the
suitable driving patterns in urban areas with heavy traffic, which requires good acceleration,
a high payload, a long driving distance, and a moderate vehicle speed.

Table 1. Targeted specification of the electric motorcycle.

Parameter Unit 1 Passenger 2 Passenger


Maximum speed km/h 80 70
Acceleration 0–30 km/h s 4 6
Driving Range at 40 km/h km 90 80
Payload kg 75 150

In order to simplify the mechanical works, the conventional engine that is directly
connected to the existing continuous variable transmission (CVT) gearbox will be replaced
by the designed SRM. The CVT amplifies and transfers torque to the rear wheel. Its gear
ratio can vary from 2.159:1 to 0.8:1, depending on the input shaft rotational speed of the
engine or motor. The centrifugal clutch begins to transmit torque to the rear wheel when
the input shaft rotates at about 1200 rpm, so the CVT is one of the key mechanical parts
taken into the calculation of the SRM’s required torque and speed characteristics. Spaces
below the footrest and inside the driver’s seat are modified to install the inverter, the battery
charger, and the 60 Ah 51.20 V Li-ion battery packs along with the battery management
system (BMS). The overall system and the developed prototype are shown in Figure 1.

Figure 1. The electric motorcycle driven by the SRM and drive system. (a) The schematic of the
electric motorcycle system; (b) the installed SRM and drive system.

The tractive force necessary to fulfill the requirements of vehicle performance in


Table 1 can be calculated using vehicle dynamic equations from [15], such as the following
key Equations:  
FT − Fa + Fg + Fw = Cm ma (1)
where the tractive force FT must exceed all resistance forces, including the aerodynamic
drag Fa , the grade resistance due to gravity Fg , and the wheel resistance Fw . The mass of
the vehicle is m, a is the acceleration, and Cm is the inertia coefficient that represents the
inertia of the rotating mass, such as the gears and wheels. Each resistance force could be
defined by
1
Fa = ρA f CD (V − Vw )2 (2)
2
Fg = mg sin α (3)
Fw = f r mg cos α (4)
where ρ is the density of air, CD is the drag coefficient of the vehicle depending mainly
on its shape, A f is the frontal area of the vehicle, Vw is the wind speed in the opposite
direction of the vehicle’s movement, and α is the grade angle. For a motorcycle, CD is
between 0.5 and 1.0.

115
Energies 2022, 15, 694

The tractive force requirements can be calculated from (1)–(4) in the design process
to achieve the desired driving performance, as shown in Table 1. Then, the traction motor
torque Tmot can be calculated by
FT rw
Tmot = (5)
Gr × ηG
where rw is the tire radius, Gr is the gear ratio, and ηG is the gearbox efficiency.
The required torque speed characteristics of the traction system in the low-speed
region include a requirement for high torque in order to shorten accelerating time, especially
during hill climbing, as well as a wide power range in the high-speed region. From the
above-mentioned equations and the dynamometer testing results of one commercial electric
vehicle, which has a similar targeted specification, it can be seen that at least 8 Nm of torque
in the low-speed region and about 7900 rpm of motor speed (which equals to a vehicle
speed of 80 km/h) are required.
Figure 2 shows the overall electric motorcycle design process under consideration.
The electric motorcycle specifications are cascading as subsequent targets for the following
subsystem design. The specification of each subsystem, namely the traction motor, the
motor drive system, battery system and mechanical structure, are taken into consideration
since they have interactive impacts on the performance of the vehicle. Several iterative
simulations have been conducted between system and subsystem levels in order to attain
the desired design goals. These interactions include motor torque–speed characteristics,
motor weight and size, DC voltage, battery DC current, drive and motor current, battery
weight and size, and battery capacity. As the battery system is one of the key components
in the electric motorcycle, the LiFePO4 battery pack has been chosen for its high capacity
per weight ratio, size, and charge/discharge rate.

Figure 2. The flowchart of the electric motorcycle design process.

116
Energies 2022, 15, 694

2.2. Traction Motor Design Consideration


The SRM was designed using a commercial analytical software package PC-SRD [16]
and a 2D finite element analysis (FEA) using JMAG software. PC-SRD was used in the
preliminary design phase to compare and adjust various design parameters, such as the
number of stator/rotor poles, the number of phases, and the winding parameters. The
software can calculate torque, current, efficiency, copper loss, iron loss, etc. It also includes
a core structure, core material data, such as BH curves, iron loss curves, winding structure,
applied voltage, rotational speed, and the on-off timing of an inverter. In order to gain high
starting torque in the low-speed region, a four-phase 8/6 pole SRM and a three-phase 6/4
pole SRM were selected for performance comparison. Moreover, the weight and cost of the
traction motor are also factors under consideration.
The design procedure of the switched reluctance machines starts with the selection
of a frame size, which is performed in accordance with the existing standard of induction
frame size (IEC71). The selection of frame size according to the IEC recommendations fixed
the outer diameter of the stator and stack length in order to maintain a similar volume.
According to the available space in the motorcycle, the key initial parameters included the
stator’s outer diameter, which was fixed to 140 mm, and its stack length, which was fixed
to 70 mm, as shown in Table 2.

Table 2. Key initial parameters of each motor.

Parameter Unit 8/6 Pole SRM 6/4 Pole SRM


Rated motor voltage (DC) V 48 48
Outer stator dimension mm 140 140
Stack length mm 70 70
Air gap mm 0.3 0.3

Based on the guideline for selecting a suitable stator and rotor pole arc in [17,18],
a stator pole arc at βs = 30 degrees and a rotor pole arc at βr = 31.5 degrees were selected as
the initial parameters for the three-phase 6/4 pole SRM. Additionally, the air gap was set
to 0.30 mm. The silicon steel with a thickness of 0.35 mm (35A300) was used and its BH
curve was used for analysis in the simulation. The magnetic characteristics of iron cores
are particularly important because SRMs do not use a permanent magnet. In this paper, the
35A300 in JIS, i.e., a general low-loss silicon steel, was used. This steel has a thickness of
0.35 mm and is equal to 36F168, 300, and V300-35A in AISI, BS, and DIN, respectively.
The torque performance in the low-speed region (2000 rpm) of the four-phase 8/6
pole and three-phase 6/4 pole SRMs was investigated for each combination of rotor yoke,
stator yoke, rotor pole width, stator pole width, as well as the number of turns and number
of strands, in order to satisfy the requirements presented in Table 1. Each parameter was
varied under the same boundaries of maximum phase current, maximum current density,
and flux density. Then, the maximum value of the developed torque was determined
(optimal torque operation). The calculated outline geometry and winding parameters of
both SRMs are shown in Figure 3 and Table 3. The preliminary results of both SRMs at
2000 rpm are shown in Table 4. It was found that both motors could generate torque larger
than 8 Nm and the three-phase 6/4 pole SRM has a 13.2% larger torque and a 14.7% larger
torque per weight than the four-phase 8/6 pole SRM. The total cost of the copper and iron
of the three-phase 6/4 pole SRM is 6.5% more expensive, and its efficiency is 9.8% lower
than the 8/6 pole SRM.

117
Energies 2022, 15, 694

Figure 3. Outline of the designed motor: (a) four-phase 8/6 pole SRM; (b) three-phase 6/4 pole SRM.

Table 3. Calculated parameter of each motor.

Parameter Unit 8/6 Pole SRM 6/4 Pole SRM


Radius to the bottom of the rotor slot (R0) mm 24 22.5
Rotor surface radius (R1) mm 32.5 32.5
Radius to the bottom of the stator slot (R2) mm 56 56
Stator outside radius (R3) mm 70 70
Air gap mm 0.3 0.3
Stator pole arc (βs ) deg 22.5 24
Rotor pole arc (βr ) deg 20 31.05
Number of turns - 9 16
Number of strands - 11 11
Wire diameter mm 0.914 0.813
Material - 35A300 35A300
Rotor inertia kg*·m2 4.8777 × 10−4 5.1453 × 10−4

Table 4. Preliminary result of the four-phase 8/6 pole and three-phase 6/4 pole SRM at 2000 rpm.

Parameter Unit 8/6 Pole SRM 6/4 Pole SRM


Phase resistance Ohm 0.012 0.032
Aligned inductance mH 0.546 2.065
Unaligned inductance mH 0.062 0.264
Theta on deg 29 34
Theta off deg 58 80
Average torque Nm 8.34 9.44
Phase current (rms) A 85.62 82.8
Shaft power W 1746.37 1976.89
Overall system efficiency % 80.46 73.27
Copper loss W 362 659
Iron loss W 55 53
Mechanical loss W 6 8
Flux density at rotor pole T 2.179 1.723
Flux density at stator pole T 1.922 2.197
Copper weight kg 1.3305 1.5959
Iron weight kg 5.3165 4.9165
Torque per weight Nm/kg 1.25 1.44
Total copper and iron cost $ 52.75 56.19

2.3. Static Performance Consideration


Figure 4 shows the torque, power, and efficiency characteristics of the designed four-
phase 8/6 pole SRM and three-phase 6/4 pole SRM at each motor rotating speed. In the
low-speed region, the motor is controlled by the current control mode, and it is controlled
by voltage control in the high-speed region. At each operating speed, the turn-on and turn-
off angles were adjusted under the same maximum value of the hysteresis band in order

118
Energies 2022, 15, 694

to search for the best combination value that serves either the optimal torque or optimal
efficient operation. To find the best turn-on and turn-off control angles for the maximum
torque and efficiency of the SRM, the grid search method, which is the easiest to implement
for the small number of parameters, was applied to reach the optimal control angles.
The optimization grid search method was implemented in Visual Basic for Applications
(VBA) of Microsoft Excel, which can be linked to PC-SRD motor design software. Grid
search was performed across different values of turn-on and turn-off angles. The different
combinations of turn-on and turn-off angles are used as input for PC-SRD to evaluate and
compare the torque and efficiency to find the optimal turn-on and turn-off angles.

Figure 4. The simulated performance of the designed four-phase 8/6 and three-phase 6/4 SRM under
optimal torque and optimal efficiency operations: (a) average torque versus rotating speed; (b) power
versus rotating speed; (c) efficiency versus rotating speed.

Two objectives were defined for the optimization of this traction motor. The first
objective is to search for the conduction angle that can give the highest torque at each
speed. The second objective is to find the conduction angle that can achieve the highest
efficiency at each speed. Constraints are that the torque of the second objective must be
within the threshold level of 1.5 Nm of the required torque, maximum DC current is 120 A,
and maximum DC voltage is 48 Vdc. Design variables are the turn-on angle (20–60 deg. for
6/4 SRM and 20–40 deg. for 8/6 SRM) and the turn-off angle (65–85 deg. for 6/4 SRM and
45–65 for 8/6 SRM). The objective function can be given in

Objective Function = η + k torque · Torque (6)

where η and k torque are the efficiency of the SRM and a factor which converts torque into an
efficiency equivalent, respectively.
Equation (6) is maximized using a grid search optimization algorithm across the
observed values of the threshold variable. Finally, the turn-on and turn-off angles for
maximum torque and maximum efficiency at each speed (1000–8000 rpm) are obtained.
For the optimal efficiency operation, the peak efficiency for the four-phase 8/6 pole SRM of
about 91.13% at 6000 rpm can be achieved, which is not different from the peak efficiency
for the three-phase 6/4 pole SRM of about 89.87% at 8000 rpm. Furthermore, for the
optimal torque operation, in the low-speed region, the maximum average torque for the
three-phase 6/4 pole SRM of about 9.57 Nm can be achieved, which is bigger than the

119
Energies 2022, 15, 694

maximum average torque for the four-phase 8/6 pole of about 8.66 Nm. That means the
three-phase 6/4 pole SRM has better torque performance in the low-speed region and good
efficiency performance in the high-speed region, which are preferable characteristics for
electric motorcycle application.
The optimal angles for maximum torque were applied to control the vehicle during
the low-speed regions or when a high dynamic response was required (braking and accel-
eration). The optimal angles for maximum efficiency were used to control the vehicle at
medium to high-speed regions or at low torque requirements to increase the range of travel
and reduce the power consumption of the battery.

2.4. Dynamic Performance Consideration


Basically, the three-phase 6/4 pole SRM has large torque ripple in the low-speed region.
Therefore, it was necessary to observe the motor phase current and torque waveforms. The
2D FEA models for dynamic simulation are revealed in Figure 5. Figures 6 and 7 show
the dynamic simulation at 2000 rpm under the optimal torque operation calculated by
PC-SRD and FEA for the four-phase 8/6 pole and three-phase 6/4 pole SRM, respectively.
The calculated values of the average torque and torque ripple rate are shown in Table 5.
According to the FEA calculation, some effects from the magnetic coupling between the
phases could be clearly observed on the phase current and torque waveforms. The results
calculated by FEA are smaller than the results calculated by PC-SRD. The three-phase 6/4
pole SRM has a larger torque ripple rate than the four-phase 8/6 pole SRM. The effect
of torque ripple would be absorbed by the CVT gearbox. However, the three-phase 6/4
pole SRM also has a larger average torque. Therefore, the three-phase 6/4 pole SRM was
selected and the final design verification was conducted. Afterwards, FEA was used to
finalize the design before constructing the prototype as shown.

Figure 5. Mesh size of FEA model (a) 6/4 SRM; (b) 8/6 SRM.

Figure 6. Four-phase current and total torque waveforms of the four-phase 8/6 pole SRM at 2000 rpm
calculated from (a) PC-SRD; (b) FEA.

120
Energies 2022, 15, 694

Figure 7. Three-phase current and total torque waveforms of the three-phase 6/4 pole SRM at 2000 rpm
calculated from (a) PC-SRD; (b) FEA.

Table 5. Calculated average torque and torque ripple rate of each motor.

8/6 Pole SRM 6/4 Pole SRM


PC-SRD FEA PC-SRD FEA
Average torque (Nm) 8.34 8.23 9.44 9.15
Torque ripple rate 0.96 0.72 1.33 1.27

The result of the preliminary three-phase 6/4 pole SRM design was verified by FEA,
which can compute the magnetic field distribution with a ferromagnetic material saturation
effect within the structure of the SRM. The results indicate that the flux density levels
are within the limit of the chosen 35A300 silicon steel material. The flux linkages of SRM
at different current levels are shown in Figure 8, where the rotor position is varied from
an unaligned to aligned position. The electromagnetic torque was calculated for various
current levels where the maximum instantaneous torque of 18 Nm was achieved at 180 A,
as shown in Figure 9.

Figure 8. Flux linkage calculation from FEA.

121
Energies 2022, 15, 694

Figure 9. Torque versus current and rotor position.

Winding arrangement is another factor that could affect the static torque performance
of the SRM. Figure 10 shows two winding configurations of the three-phase 6/4 pole SRM
when a single phase (phase A) is excited with a unipolar operation. Figure 11 shows the
flux density and flux path of the short-flux path and long-flux path winding configurations,
respectively. The comparison of the static torque developed at each rotor position for
both the short-flux and long-flux paths of the three-phase 6/4 pole SRM is based on
the same conduction loss condition, which is also shown in Figure 12. The long-flux
path winding could produce more positive torque during the effective torque production
region, which is 45 mechanical degrees for the three-phase 6/4 pole SRM. The short-
flux path winding could produce lower peak torque and it produces negative torque at
77–90 degrees. The three-phase 6/4 pole SRM should also be excited by bipolar excitation
with a sinusoidal waveform [19] in order to reduce the torque ripple rate. However, the
average torque and efficiency are also smaller than the unipolar excitation. Therefore, the
static torque performance of the three-phase 6/4 pole SRM excited by bipolar excitation is
not investigated here.

Figure 10. Winding configuration of the three-phase 6/4 pole SRM: (a) short-flux path; (b) long-flux path.

122
Energies 2022, 15, 694

Figure 11. Flux distribution of each winding configuration: (a) short-flux path; (b) long-flux path.

Figure 12. Static phase torque for each winding configuration.

3. Construction of the Motor and Drive System


Figure 13 shows the constructed three-phase 6/4 pole SRM prototype and drive
system. In total, 30 units of the three-phase 6/4 pole SRM were manufactured by the
private company in order to control the quality of the prototypes.

Figure 13. The prototypes of the three-phase 6/4 pole SRM and drive system.

The motor drive system employs the conventional three-phase, asymmetric half-
bridge inverter. This topology has the advantages of simplicity, robustness, and fault
tolerance, where each phase can be controlled independently with unidirectional current
flow. These greatly enhance the safety features of the electric motorcycle drive system.
The control block diagram for the SRM drive system is shown in Figure 14, where the
user turns the accelerator to define the torque command as an input. The motor torque is
estimated based on the torque to current block. The optimal turn-on angle, turn-off angle,
and current command for maximum efficiency operation (optimal efficiency operation)
at each torque and speed level will be calculated off-line by FEA and put into the 3D

123
Energies 2022, 15, 694

look-up table, as shown in Figure 15. At each torque command and detected motor
speed, the corresponding current command and on/off angles will be calculated by the
following Equation.
   
s − s2 τi − τ2 τ − τ1 s − s1 τi − τ2 τ − τ1
f (τi , si ) = i f (τ1 , s1 ) + i f (τ2 , s1 ) + i f (τ1 , s2 ) + i f (τ2 , s2 ) (7)
s1 − s2 τ1 − τ2 τ2 − τ1 s2 − s1 τ1 − τ2 τ2 − τ1
where τi and si are the torque command and detected motor speed, respectively.

Figure 14. The control block diagram.

Figure 15. Optimal value of (a) current command; (b) turn-on angle; and (c) turn-off angle at each
torque command and motor speed level for optimal efficiency operation.

The rotor position is also a necessary parameter for the indirect torque controller and
current controller in the drive system. For the electric motorcycle application, especially
when the vehicle with maximum payload starts from the standstill condition on the slope,
the initial rotor position is very necessary for high starting torque development. Some
low-cost, incremental optical encoders may be used for rotor position detection. However,
they cannot detect the initial rotor position at a standstill condition, leading to torque
jerk [20], which seriously affects the safety of the electric motorcycle. Furthermore, due
to the harsh environment on the road, with dust, water, and vibration, a more reliable
detection system for a non-contact rotor position is required. In this paper, the contactless
magnetic-based position sensor was installed on the rotor shaft for position and speed
feedback, as shown in Figure 16. It is a very low-cost and high-reliability solution. With the
proposed mechanical coupling device using a bolted shaft and ball bearing, the effect of
the deviated distance between the permanent magnet and sensing circuit caused by speed
variation and vibration could be eliminated effectively. A resolution of the rotor position
detection of about 0.87 degrees could be achieved from 0 to 8000 rpm. The proposed initial
rotor position estimation method [20] based on pulse injection and phase current profile has
a maximum error of about 2.5 degrees, but it could identify the sector of the rotor position.
Therefore, the combination of the proposed rotor position estimation and magnetic sensing
system could increase the reliability of the electric motorcycle.

124
Energies 2022, 15, 694

Figure 16. Rotor position detection system: (a) mechanical coupling device; (b) a constructed rotor
position detection system installed on the three-phase 6/4 pole SRM.

The motor phase current is sensed by current sensor model ACS758ECB-200U which
can be mounted on a PCB. The current regulation is also performed by an inner digital
hysteresis controller. A 32-bit micro-controller, STM32F103RET6, and a power MOSFET,
IRF4468, were used. Figure 17 shows the battery pack and circuit board in the bat-
tery management system. In one battery pack, a total of eight cells of LiFePO4 model
NLC36130185PF (3.2 V 63 Ah) were connected in a series and two sets of battery packs
were installed in each electric motorcycle. The voltage of each cell is monitored by the
cell monitoring board and it is transmitted to the master board. The master board will
perform passive balancing by bypassing the current to the cell that has the lower capacity
and it will make a decision for communication to the quick charger and interlock circuit for
safe operation.

Figure 17. (a) Lithium ion battery pack; (b) master board of the battery management system; (c) cell
monitoring board of the battery management system.

4. Experiment Results
4.1. Motor and Drive System Test Result
The static test of the SRM prototype was conducted. The electrical characteristics of
the prototype, such as the resistance and inductance, can be measured via a static test with
the locked rotor. Table 6 shows the measured inductance and simulated value obtained
from FEA. The small discrepancy may be caused by the effect of 2D FEA.

Table 6. Measured inductance value compared to FEA.

Phase L-Aligned (mH) L-Unaligned (mH) L-Ratio


A 1.9767 0.2960 6.6774
B 1.8016 0.3342 5.3904
C 1.9385 0.2864 6.7666
FEA 2.0652 0.2643 7.8150

Figure 18 shows the efficiency map of the SRM prototype from the test results on the
test bench. With the limited capacity of the testing equipment, the SRM drive was supplied
with a 50 VDC voltage source and maximum DC current of about 100 A, which is less than
the value used in the simulation. Therefore, performance of up to 7000 rpm was measured.

125
Energies 2022, 15, 694

Figure 18. Efficiency map of the constructed three-phase 6/4 pole SRM.

The maximum torque measured via a torque transducer was approximately 8.6 Nm
at 3000 rpm. The motor has an efficiency of more than 80% at a high speed and medium
torque output range (81.85% at 7000 rpm), which is very suitable for an electric motorcycle
running at medium to high speed. This load condition is the most practical scenario for the
normal usage of a general purpose electric motorcycle.

4.2. Electric Motorcycle Performance Test Result


The performance of the constructed electric motorcycle was tested by a chassis dy-
namometer, as shown in Figure 19. The driving range of the electric motorcycle at a constant
40 km/h cruising speed (motor speed 4686 rpm) with a 150 kg pay load was measured.
Figure 20 shows the measured values during the test run. From the starting time to 140 min,
the measured current was about 30 A and the measured voltage gradually decreased from
51.20 V. When the running time reached 140 min, the battery voltage was lower than the bat-
tery threshold, and the drive system was then turned off. The total driving range of about
98 km could be achieved. The maximum vehicle speed and acceleration from 0 to 30 km/h
during on-road testing with full payload (150 kg) were measured and compared to the
commercial electric motorcycle, as shown in Figure 21. The maximum speed could reach
82 km/h and the acceleration time was about 5.1 s, which is better than the requirements
in Table 1.

Figure 19. Test run of the constructed electric motorcycle on a chassis dynamometer.

126
Energies 2022, 15, 694

Figure 20. (a) Measured DC voltage and distance; (b) measured DC current and vehicle speed.

Figure 21. (a) Maximum vehicle speed; (b) acceleration from 0 to 30 km/h.

The constructed electric motorcycle was tested and complied with the electromagnetic
compatibility standards, including radiated emission and immunity testing based on
CISPR25:2002 and ISO 11452-2, respectively, as shown in Figure 22. Furthermore, the
SRM drive system was also in compliance with the IEC1-60335 standard for the safety of
household and similar electrical appliances. The SRM was also tested and complied with
the IEC60529:2001 (IP56) standard. All 30 units of the completed electric motorcycles were
produced and transferred to the Provincial Electricity Authority (PEA) for their internal
routine services.

Figure 22. EMI/EMC testing of the constructed electric motorcycle.

5. Conclusions
This paper described the performance improvement of an electric motorcycle driven
by an SRM and drive system. The static performance comparison of the three-phase 6/4
pole and four-phase 8/6 pole SRM was undertaken. It was found that the three-phase
6/4 pole SRM had a larger starting torque and torque per weight in the low-speed region.
In the high-speed region, both SRMs could reach about 90% with the optimal efficiency
operation. The average and torque ripple rate of both SRMs were also investigated by
dynamic simulation. The torque ripple rate of the three-phase 6/4 pole SRM was a bit
larger than the four-phase 8/6 pole SRM. However, the three-phase 6/4 pole SRM also
had a larger average torque. Therefore, the three-phase 6/4 pole SRM was also suitable

127
Energies 2022, 15, 694

for the propulsion system of the electric motorcycle. The performance of the three-phase
6/4 SRM was also investigated by FEA. The implementation of indirect torque control and
rotor position detection in the drive system was proposed. According to the testing results
obtained on the test bench, the highest motor efficiency of about 82% could be achieved.
The vehicle performance, including a maximum speed of 82 km/h and a driving range of
98 km at a constant speed condition of 40 km/h, could be also achieved.

Author Contributions: Conceptualization, S.K., P.J., P.S., R.P. and K.T.; methodology, S.K., P.J., P.S.
and K.T.; validation, R.P., S.K., P.J., P.S. and K.T.; formal analysis, S.K., P.J., P.S., R.P. and K.T.;
investigation, R.P., P.S. and K.T.; resources, S.K., P.J. and P.S.; writing—original draft preparation, K.T.;
writing—review and editing, R.P., S.K., P.J., P.S. and K.T.; visualization, R.P. and S.K.; supervision, P.S.
and K.T. All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by the Provincial Electricity Authority Research Fund of Thailand.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Kerdlap, P.; Gheewala, S.H. Electric motorcycles in Thailand—A life cycle perspective. J. Ind. Ecol. 2016, 20, 1399–1411. [CrossRef]
2. Bostanci, E.; Moallem, M.; Parsapour, A.; Fahimi, B. Opportunities and challenges of switched reluctance motor drives for electric
propulsion: A comparative study. IEEE Trans. Transp. Electrif. 2017, 3, 58–75. [CrossRef]
3. Rahman, K.M.; Fahimi, B.; Suresh, G.; Rajarathnam, A.V.; Ehsani, M. Advantages of switched reluctance motor applications to EV
and HEV: Design and control issues. IEEE Tran. Ind. Appl. 2000, 36, 111–121. [CrossRef]
4. Lan, Y.; Benomar, Y.; Deepak, K.; Aksoz, A.; Baghdadi, M.; Bostanci, E.; Hegazy, O. Switched reluctance motors and drive systems
for electric vehicle powertrains: State of the art analysis and future trends. Energies 2021, 14, 2079. [CrossRef]
5. Jiang, J.W.; Bilgin, B.; Emadi, A. Three-phase 24/16 Switched Reluctance Machines for a Hybrid Electric Powertrain. IEEE Trans.
Transp. Electrif. 2017, 3, 76–85. [CrossRef]
6. Kiyota, K.; Chiba, A. Design of switched reluctance motor competitive to 60-kW IPMSM in third-generation hybrid electric
vehicle. IEEE Trans. Ind. Appl. 2012, 48, 2303–2309. [CrossRef]
7. Wang, S.; Zhan, Q.; Ma, Z.; Zhou, L. Implementation of a 50-kW four-phase switched reluctance motor drive system for hybrid
electric vehicle. IEEE Trans. Magn. 2005, 41, 501–504. [CrossRef]
8. Lin, J.; Schofield, N.; Emadi, A. External rotor 6-10 switched reluctance motor for an electric bicycle. IEEE Trans. Transp. Electrif.
2015, 1, 348–356. [CrossRef]
9. Terzic, M.V.; Bilgin, B.; Emadi, A. Switched reluctance motor design for a forklift traction application. In Proceedings of the 2018
XIII International Conference on Electrical Machines (ICEM), Alexandroupoli, Greece, 3–6 September 2018; pp. 812–818.
10. Andrada, P.; Blanque, B.; Capo, M.; Gross, G.; Montesinos, D. Switched reluctance motor for light electric vehicles.
In Proceedings of the 2018 20th European Conference on Power Electronics and Applications (EPE’18 ECCE Europe), Riga, Latvia,
17–21 September 2018; pp. 1383–1388.
11. Vosswinkel, M.; Lohner, A.; Platte, V.; Hirche, T. Design, production, and verification of a switched-reluctance wheel hub drive
train for battery electric vehicles. World Electr. Veh. J. 2019, 10, 82. [CrossRef]
12. Lin, C.H. Switched reluctance motor drive for electric motorcycle using HFNN controller. In Proceedings of the 2007 7th
International Conference on Power Electronics and Drive Systems, Bangkok, Thailand, 27–30 November 2007; pp. 1383–1388.
13. Tomczewski, K.; Wrobel, K.; Rataj, D.; Trzmiel, G. A switched reluctance motor drive controller based on an FPGA device with a
complex PID regulator. Energies 2021, 14, 1423. [CrossRef]
14. Anvari, B.; Toliyat, H.A.; Fahimi, B. Simultaneous optimization of geometric and firing angles for in-wheel switched reluctance
motor drive. IEEE Trans. Transp. Electrif. 2018, 4, 322–329. [CrossRef]
15. Ehsani, M.; Gao, Y.; Longo, S.; Ebrahimi, K.M. Modern Electric, Hybrid Electric, and Fuel Cell Vehicles Fundamentals, Theory, and
Design, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2010.
16. SPEED Consortium University of Glasgow. User’s Manual: PC-SRD; Version 8.8, CD-Adapco; SPEED Consortium University of
Glasgow: Glasgow, UK, 2011.
17. Vijayraphavan, P. Design of Switched Reluctance Motors and Development of a Universal Controller for Switched Reluctance
and Permanent Magnet Brushless DC Motor Drives. Ph.D. Thesis, Virginia Polytechnic Institute and State University, Blacksburg,
VA, USA, 2001.
18. Miller, T.J.E. Switched Reluctance Motors and Their Control; Magna Physics Publishing and Clarendon Press: Oxford, UK, 1993.

128
Energies 2022, 15, 694

19. Liu, X.; Zhu, Z.Q.; Hasegawa, M.; Pride, A.; Deodhar, R.; Maruyama, T.; Chen, Z. Performance comparison between unipolar and
bipolar excitations in switched reluctance machine with sinusoidal and rectangular waveforms. In Proceedings of the 2011 IEEE
Energy Conversion Congress and Exposition, Phoenix, AZ, USA, 17–22 September 2011; pp. 1590–1595. [CrossRef]
20. Tungpimolrut, K.; Jitkreeyarn, P.; Kachapornkul, S.; Somsiri, P.; Chiba, A. Initial rotor position estimation of a SRM drive installed
in an electric vehicle. IEEJ Trans. Electr. Electron. Eng. 2011, 6, 594–600. [CrossRef]

129
energies
Article
Design and Thermal Analysis of Linear Hybrid Excited Flux
Switching Machine Using Ferrite Magnets
Himayat Ullah Jan 1, *, Faisal Khan 1 , Basharat Ullah1, *, Muhammad Qasim 1 , Ahmad H. Milyani 2,3
and Abdullah Ahmed Azhari 4

1 Department of Electrical and Computer Engineering, COMSATS University Islamabad,


Abbottabad 22060, Pakistan; [email protected] (F.K.); [email protected] (M.Q.)
2 Department of Electrical and Computer Engineering, King Abdulaziz University, Jeddah 21589, Saudi Arabia;
[email protected]
3 Center of Research Excellence in Renewable Energy and Power Systems, King Abdulaziz University,
Jeddah 21589, Saudi Arabia
4 Computer Information Technology, The Applied College, King Abdulaziz University,
Jeddah 21589, Saudi Arabia; [email protected]
* Correspondence: [email protected] (H.U.J.); [email protected] (B.U.)

Abstract: This paper presents a novel linear hybrid excited flux switching permanent magnet machine
(LHEFSPMM) with a crooked tooth modular stator. Conventional stators are made up of a pure iron
core, which results in high manufacturing costs and increased iron core losses. Using a modular
stator lowers the iron volume by up to 18% compared to a conventional stator, which minimizes
the core losses and reduces the machine’s overall cost. A crooked angle is introduced to improve
the flux linkage between the stator pole and the mover slot. Ferrite magnets are used with parallel
magnetization to reduce the cost of the machine. Two-dimensional FEA is performed to analyze and
Citation: Jan, H.U.; Khan, F.; Ullah, evaluate various performance parameters of the proposed machine. Geometric optimization is used
B.; Qasim, M.; Milyani, A.H.; Azhari,
to optimize the split ratio (S.R) and winding slot area (Slot area ). Genetic algorithm (GA) is applied
A.A. Design and Thermal Analysis of
and is used to optimize stator tooth width (STW ), space between the modules (SS), crooked angle (α),
Linear Hybrid Excited Flux
and starting angle (θ). The proposed model has a high thrust density (306.61 kN/m3 ), lower detent
Switching Machine Using Ferrite
force (8.4 N), and a simpler design with higher efficiency (86%). The linear modular structure makes
Magnets. Energies 2022, 15, 5275.
https://doi.org/10.3390/
it a good candidate for railway transportation and electric trains. Thermal analysis of the machine
en15145275 is performed by FEA and then the results are validated by an LPMEC model. Overall, a very good
agreement is observed between both the analyses, and relative percentage error of less than 3% is
Academic Editors: Vladimir Prakht,
achieved, which is considerable since the FEA is in 3D while 2D temperature flow is considered in
Mohamed N. Ibrahim and Vadim
the LPMEC model.
Kazakbaev

Received: 30 June 2022 Keywords: linear machine; flux switching machine; modular stator; crooked tooth; ferrite magnet;
Accepted: 17 July 2022 genetic algorithm; thermal analysis; LPMEC model
Published: 21 July 2022

Publisher’s Note: MDPI stays neutral


with regard to jurisdictional claims in
published maps and institutional affil- 1. Introduction
iations. Increase in the industrialization of the modern world increases the pollution caused
by automobiles and other transportation sources. The modern world is shifting toward
more efficient and pollution-free hybrid electric vehicles because of the environmental
issues and to escape an imminent threat of energy scarcity. Railways also are the main
Copyright: © 2022 by the authors.
constituent of the transportation system. Major intercity transportation is carried out by
Licensee MDPI, Basel, Switzerland.
trains. Rotary motors have been used previously in the literature for the rail infrastructure,
This article is an open access article
but the gearing system used to convert rotary motion to linear lowers the overall efficiency
distributed under the terms and
of the system. The trend is now shifting toward linear motors as they can be directly used
conditions of the Creative Commons
Attribution (CC BY) license (https://
because of their direct nonadhesive thrust force without using the gearing system.
creativecommons.org/licenses/by/
Linear induction motors (LIMs) have been used recently for the transportation system.
4.0/). Primarily, the LIMs need a lower value of volume than most of the traditional motors,

Energies 2022, 15, 5275. https://doi.org/10.3390/en15145275 131 https://www.mdpi.com/journal/energies


Energies 2022, 15, 5275

reducing the cross-sectional area of tunnels. A single-sided LIM is analyzed and studied
in [1] with the derivation of the equivalent circuit and its analysis. Control strategies
were devised for the LIMs in [2,3]. A modified model-predictive control was proposed for
LIMs, and the results were analyzed. Effects of different secondaries were studied and
investigated in [4,5]. Considering all these design optimizations and control algorithms,
the main problems faced by LIMs are having lower efficiency and low power factor with
respect to eddy currents, copper losses, and effects of edging, resulting in a high system
and maintenance cost.
In contrast, linear permanent magnet (LPM) motors have been proven to have high
efficiency, high power density, and high power factor. LPMs have many advantages, but
they still have the drawback of using a large number of magnets. Mostly, the magnets
are placed on the long stator, which not only makes it complex but also makes it costly.
The linear flux switching machine (LFSM) is extensively studied nowadays because of its
PM placement in the short mover and having a simple iron stator. In addition to these
advantages of PM machines, LFSPM machines have several other advantages, such as
easy maintenance, easy heat management system, and lower cost of the secondary, which
makes it unique for purposes such as trains and railway stations [6,7]. However, LFSPM
suffers from high detent force due to slot and end effects. Slot effect can be suppressed by
adjusting the length and width of the PM [8]. It can be used to reduce detent force, but it
has a negative effect on the thrust force, reducing it considerably. In [9,10], staggered tooth
and semi-closed slots were proposed for reducing detent force and end effect, but it has the
drawback of making the winding arrangement difficult and making the whole machine
complex. To reduce the end effect in linear machines, auxiliary poles (APs) were proposed
in [11,12], but they also increase the total harmonic distortions (THDs) of the back-EMF
profile. Modular structure was used in [13,14] to reduce the end effect caused by unstable
magnetic circuits.
Thermal environment greatly affects the working conditions of a machine. Authors
in [15] used the Arrhenius model along with a coaxial multi-slot antenna. Authors in [16]
analyzed various deformations along with the wear and tear because of the temperature rise.
Ref. [17] only considered the distribution of temperature while ignoring the temperature
rise with the passage of time. Refs. [18,19] divided the whole machine into various different
parts and analyzed temperature rise in each of the parts, which increased the efficiency
of the method. Authors in [20,21] used the same method of dividing the machine various
isotropic parts and analyzing the temperature in each part and then comparing the average
temperature rise in the whole of the machine.
This paper proposes a novel LHFSPM machine with crooked tooth modular stator.
Two DC excitation sources are placed below and above a ferrite magnet, and overlapped
concentrated winding is used for the armature. The stator of the machine is in the form
of a U-shaped module with a crooked angle. The whole model is designed and analyzed
in JMAG, registered version 20.1. No-load and loaded studies of the machine are carried
out, and optimization techniques were used to improve the thrust force characteristics.
This paper is divided into six further sections: Section 2 discusses the structure, design,
and working principle of the machine; Section 3 covers optimization techniques and
their effect on the performance of the machine; Section 4 discusses the electromagnetic
performance evaluation using a number of study techniques to reach the proposed model;
Section 5 discusses the thermal analysis of the machine and comparison between the FEA
thermal analysis and LPMEC model; Section 6 presents the comparison of proposed and
conventional machine; Section 7 provides the conclusion, which is an overview of the
whole paper, and all the necessary points are discussed.

2. Design and Working Principle


The design of the proposed model is shown in Figure 1. DC field windings (shown by
cyan color) and armature windings (shown by green color) are placed on the short mover.
Permanent magnets (PMs) (shown by magenta color) are placed inset in the mover poles

132
Energies 2022, 15, 5275

magnetized in a parallel direction. The stator of the model is in the form of U-shaped
modules, which are placed at a certain distance from each other, having no direct electrical
or magnetic contact. Such an arrangement uses less iron than the conventional stator
and not only decreases the cost of the machine but also improves the overall efficiency of
the machine. A conventional stator is made up of a full-length iron core which becomes
impractical for countries such as Pakistan and other developing countries.

Figure 1. 2D view of the proposed model.

To minimize the detent force of the machine, a suitable selection of mover slot Ms and
stator pole S p is made using Equation (1) [13].
 
1±n
S p = Ms (1)
2q

where q denotes the number of phases and n represents any natural number. For various
values of n, different stator pole numbers are studied and analyzed. It was noted that the ma-
chine has a sinusoidal flux linkage and unidirectional thrust force when Ms /S p = 6/5, 6/7.
The average thrust force for 6/5 is higher than 6/7, so it is considered for further analysis.
The velocity of the machine is dependent on the input source frequency and pole pitch of
the machine and can be found by using Equation (2) [22].

v = f τs (2)

where v is the velocity, f is the frequency of the source, and τs is the pole pitch of the
machine. Flux linkage of the machine greatly varies (periodically) with the position of
the mover relative to the stator position since the flux path changes as the mover changes
its position. Considering phase A at two different points, both maximum and minimum
flux linkage are analyzed, as shown in Figure 2a,b, respectively. At point 1, flux linkage
due to phase A is at a positive maximum as the mover pole and stator tooth is completely
aligned. Flux flows through the magnet, mover pole, air gap, and then into the stator tooth,
completing the flux path back into the mover. A point 2, the flux linkage of phase A is at a
negative maximum as the mover pole stator tooth is completely misaligned, while phase B
and phase C have some value. The three phases are completely (120 degrees) apart. The
direction of PM magnetization is set horizontal to that of the primary moving direction.
The direction of PM and armature winding and DC is either clockwise or anticlockwise,
to strengthen the overall machine’s flux linkage. Parameters of the machine are defined
in Figure 3. The details of the parameters and their numerical values are tabularized in
Table 1.

(a) (b)
Figure 2. Flux linkage at two different positions. (a) Point 1; (b) Point 2.

133
Energies 2022, 15, 5275

Table 1. Detailed parameters of the proposed model.

Parameters Values
Pole Pitch, τs 26.1984 mm
Velocity, V 4 m/s
Stator Height, Hs 12.56 mm
Stack Length 90 mm
Mover Height, Hm 12.56 mm
Stator Tooth Width, STW 8.33338 mm
Air gap, L g 0.8 mm
Armature winding turns 136
Maximum current DC 5.76 A
Permanent magnet Ferrite_Br = 0.4 T
Maximum thrust force 150.1823 N
Split ratio, S.R 0.246
Armature slot width, Aw 2.4 mm
Armature slot length, A L 31.241 mm
Upper DC slot width, DCUW 12.55 mm
Upper DC slot length, DCUL 4.3 mm
Lower DC slot width, DCLW 3.119 mm
Lower DC slot length, DCLL 9.264 mm

Figure 3. Parameters of the proposed model.

A flux density nephogram of the machine is shown in Figure 4, which reflects the
magnetic flux density at various points of the machine, the path followed by the flux, and
how flux switches from one mover pole to another pole through the modular stator. Slight
saturation can be seen in the red regions in the nephogram, but it cannot be considered
to cause a heating effect in the machine. The maximum flux density in the mover is 1.8 T,
while in the case of the stator, maximum flux density is 1.42 T.
Coil configuration is dependent on the number of mover slots. Once the mover slot
combination is confirmed, the coil configuration can be adjusted accordingly. Coil span is

134
Energies 2022, 15, 5275

the axial distance through which a coil is wound. The coil span of the machine depends on
the type of winding configuration used, either concentrated or simple.

max [ f ix [ M
Ts ]]
s
1
Cs = (3)
f or concentrated winding 1

where Ms represents number of mover slot, S p represents number of stator tooth, and f ix
function is used to return integer value only. In case of a three-phase balanced system, the
phases are separated 120 degrees apart. The phase separation for a machine can be found
by C0 .
Nm
C0 = (1 + 3k) (4)
3Ns
k is any integer value, e.g., 0, 1, 2, 3, . . . N2s . If a suitable value for k is not found, then
that value of slot and pole combination shall not be chosen.

Figure 4. Nephogram of machine flux density.

3. Optimization and Refinement of Machine Parameters


To increase the thrust force and efficiency of the machine, different machine parameters
were optimized using single variable geometric optimization and JMAG inbuilt optimiza-
tion (genetic algorithm (GA)). Thrust force amplification is considered the main target. In
geometric optimization, a series of consecutive values are considered for geometry, and
then the resulting thrust force is analyzed.

3.1. Geometric Optimization


Leading parameters such as split ratio (S.R) and armature slot width (Aw ) were
optimized using geometric optimization. Figure 5 shows the flowchart of single variable
geometric optimization.
Determination of optimal split ratio is a very important process of designing a machine
as it decides not only the average thrust force but also the overall cost of the machine. If the
selected value of the split ratio is low, it will reduce the stator height, which is suitable for
railway transits. On the other hand, it will increase mover height and mass, resulting in the
reduction of the average thrust force. The higher value of the split ratio resolves the mover
mass problem but increases the stator volume, making the machine costly. The split ratio
of the machine can be found by Equation (5) [23]. Table 2 shows the performance of the
machine at different values of S.R.
Hs + L g
S.R = (5)
Hs + Hm + L g

Slot area
Aw = (6)
AL

135
Energies 2022, 15, 5275

Table 2. Performance indicator.

S.R 0.15 0.225 0.235 0.246 0.257


TFaverage (N) 119.17 121.72 128.43 133.48 129.13
ϕ( p− p) (Wb) 0.010 0.0107 0.017 0.012 0.018
Fd( p− p) (N) 0.0237 0.215 0.281 0.301 0.260
TFd (kN/m3 ) 243.30 248.51 262.21 272.52 263.64
THD (%) 12.93 11.23 13.75 14.1 13.42

Figure 5. Flowchart of geometric optimization.

The suitable height and width of the slot area are selected by keeping the overall slot
area constant and changing the width and height of the machine. Increasing the width
tends to decrease the height of the slot, and decreasing the width increases the height of the
slot. The height and width of the machine are interrelated by Equation (6). Table 3 shows
the thrust force profile at different slot width and height values.

Table 3. Performance after geometric optimization.

Indicator Stage1 Stage2 Stage3 Stage4 Stage5 Stage6


A L (mm) 35.71 34.09 32.60 31.25 28.84 27.77
Aw (mm) 2.1 2.2 2.3 2.4 2.5 2.6
Slot area (mm2 ) 75 75 75 75 75 75
TFaverage (N) 131.21 132.8 133.48 134.17 135.01 135.98

136
Energies 2022, 15, 5275

3.2. Genetic Optimization


Stator tooth width, stator module spacing, crooked angle of the tooth, and starting
angle of the machine are the four parameters that are optimized using genetic optimization.
The width of the tooth helps with better flux linkage and better alignment of the mover
pole and stator tooth. Teeth provide the necessary path for the flux linkage, and if teeth are
too thin, the machine will suffer from saturation, and if the teeth are too thick, the flux will
not switch into the next mover pole. GA was used to select a better-suited stator width,
resulting in a higher thrust force and higher flux linkage.
The stator of the proposed machine is modular, and the modules are spaced at a
certain distance, so better placement of the module becomes very important. It not only
increases the coil flux linkage but is also used to minimize the usage of iron and the cost of
the machine. GA single variable optimization is used to select the optimal spacing distance
between the modules. Complete details of initial values, final values, and the constraint for
GA are given in Table 4.

Table 4. Performance of the machine after Genetic Optimization.

Parameters Initial Value Final Value Constraints TF


STW 12.24 mm 8.338 mm 7 mm < STW < 13 mm 136.08 N
SS 11 mm 13.01 mm 10 mm < SS < 15 mm 140.601 N

The crooked angle of the stator tooth is analyzed, and its effect on the machine’s
performance is evaluated. It is the angle made by the inner side of the tooth with the
yoke of the stator. Genetic optimization is employed, and it was witnessed that initially
when the angle α is increased from 0◦ to 6.22◦ , the performance of the machine increases
considerably; the best performance being noted at 6.22◦ . For an angle between 6.22◦ to 20◦ ,
a minute decrease is experienced in the thrust force. Figure 6 shows the crooked angle of
the stator tooth, while Figure 7 shows the trend of how thrust force varies with the angle
variation.

Figure 6. Crooked angle.

Figure 7. Variation of thrust force with crooked angle.

137
Energies 2022, 15, 5275

Figure 8 shows the average thrust force at different armature current starting angles.
Three different Je are considered, and the resultant thrust force is analyzed. It can be seen
that the machine performs best at a 0◦ angle and the thrust force keeps on decreasing,
going away from the origin in either direction. GA is used to optimize the starting angle,
and its effect on the average thrust force is analyzed. The comparison of before and after
optimization flux linkage and thrust force is shown in Figures 9 and 10, respectively.
Table 5 shows before and after optimization values of parameters such as detent force,
thrust force, and THD.

Figure 8. Thrust force at different starting angle.

Figure 9. Flux linkage before and after optimization.

Figure 10. Thrust force comparison before and after optimization.

138
Energies 2022, 15, 5275

Table 5. Overall performance of the machine.

Items Initial Value After Optimization Thrust Force Before Thrust Force After THD F(detent p− p)
S.R 0.225 0.246 129.5353 N 133.4818 N 14.1% 0.2601 N
Aw 2.3 mm 2.6 mm 133.4818 N 135.9811 N 13.7% 0.498 N
STW 12.24 mm 8.33 mm 135.9811 N 136.08 N 11.79% 2.1 N
SS 11 mm 13.01 mm 136.08 N 140.60 N 13.21% 4.12 N
α 0 6.22 140.60 N 148.608 N 12.01% 7.2 N
θ 10 0 148.608 N 150.1823 N 12.54% 8.4 N

4. Analysis of Electromagnetic Performance


Parameters such as no-load flux linkage, detent force, total harmonic distortion (THD)
of U phase, and thrust force are investigated for a boundary period of 1. No-load flux
linkage ϕ( p− p) , detent force (Fd( p− p) ), and thrust force are directly calculated from the FEA.
Mathematical calculations were performed to find TFd for on-load and THDs of no-load
flux linkage. Fourier transform of no-load flux linkage is taken, after which Equation (7) is
used to find THDs.
Σkk=2 ϕ2k
THD = (7)
ϕ1
where ϕ1 represents fundamental component, and ϕ2 to ϕk are the harmonics. Thrust force
density with respect to mover volume of on-load study is calculated by Equation (8).

TFaverage
TFd = (8)
Vm

Figure 11 represent the three-phase no-load flux linkage, all the phases are purely
sinusoidal. A difference can be noted between the positive maximum value and the
negative value, which points towards the presence of a leakage flux. The flux regulation
capability at various DC excitation current is shown in Figure 12. The figure shows that the
flux of the proposed machine can be easily controlled by varying the DC current.

Figure 11. No-load flux linkage.

139
Energies 2022, 15, 5275

Figure 12. Flux regulation capability of the proposed machine.

Detent force is analyzed in the no-load study when the AC circuit is open, and only
DC is fed to the machine. The presence of a magnet makes it the combination of both DC
and PM. Depending on its value, it either pulls the machine backward or pushes it forward.
Positive detent force helps push the machine forward, while negative detent force pulls the
machine backward. This push and pull are the main reasons for thrust force ripple [24,25].
A bipolar detent force can be seen in Figure 13.
The thrust force of the machine is unipolar in nature, as shown in Figure 10. The effect
of detent force pull and push can be observed from the thrust force graph. When the value
of detent force decreases, i.e., at angle 30◦ to 120◦ , 160◦ to 240◦ , and 300◦ to 360◦ , the thrust
force decreases, but when the detent force increases, the thrust force also increases. Table 6
shows the detailed values for THD, peak-to-peak no-load flux linkage, thrust force, and
thrust force density.
Performance of the machine is evaluated at different Ja ; it can be observed that the
thrust force increases linearly up to some extent, but then the linearity is disturbed because
the machine is moving toward saturation.
Figure 14 shows thrust force at different Ja and Je values. Two different values for Je
are considered, and the respective thrust force is shown. Thrust force profiles and power of
the machine are evaluated at different velocities and are presented in Figures 15 and 16.

Figure 13. Detent force.

140
Energies 2022, 15, 5275

Table 6. Performance indicators.

Parameters Values
ϕ( p− p) 0.12 T
ϕ(+max) 0.05 T
ϕ(−max) 0.07 T
THD (%) 12.10
TFaverage 129.53 N
F(detentp− p) 8.4 N
TFd 263.64 kN/m3

Figure 14. Thrust force at different Ja .

Figure 15. Thrust force at different velocities.

Figure 16. Output power at different velocities.

141
Energies 2022, 15, 5275

Different points are considered in the thrust force and velocity graph, and both iron
core losses and copper losses are calculated. Overall efficiency of the machine is calculated
considering both iron and copper loss. Copper loss of the machine can be calculated using
Equation (9).
Pc = IρJLNQ(1000) (9)
where I represents armature current, ρ represents resistivity of the conductor, L is the length
of the wire, Ja is the current density of the wire, N is the number of conductors, and Q
denotes number of slot pairs.
Iron losses of the machine are calculated directly from JMAG simulation, and the overall
efficiency of the machine is calculated. Figure 17 shows losses and overall efficiency of the
machine at different velocities. Use of a modular stator reduces the iron volume, thus lowers
the iron core losses, increasing the efficiency of the machine. A total of 13% improvement in
efficiency is noted using a modular stator because of the iron losses minimization.

Figure 17. Efficiency at different points.

5. Thermal Analysis
Thermal analysis of the machine has three broad steps: the first one is calculation
of three dimensional magnetic losses, thermal analysis, and comparing analytically with
lumped parametric model equivalent circuit (LPMEC). The whole process of thermal
analysis is shown in the form of a flowchart in Figure 18. The whole procedure of analysis
and model building is defined by [20].

5.1. Magnetic Analysis


Since accuracy of the machine is higher in the 3D model, 3D magnetic analysis is
considered. Losses calculated are calculated after necessary condition and material setting,
and then the losses are used in the thermal analysis.

5.2. Thermal Analysis


Since thermal properties of the machine greatly depend on the nature of the material,
conductivity of each material is calculated along with the specific heat capacity of the
materials. Various properties of materials are tabulated in Table 7. Room temperature
is considered as a reference for the whole process. Three different types of boundaries
were defined for various contact types. For direct contact between the two parts, a contact
thermal resistance boundary is defined, and heat transfer boundaries are defined for points
where heat transfer in the form of convection takes place. Radiation type of heat transfer is
ignored in this study. Figure 19 represents the distribution of temperature throughout the
machine design. Most of the heat accumulation takes place in the mover as it contains all
the active parts (coils and magnets). The stator remains at room temperature as there is no
direct contact between the mover and the stator.

142
Energies 2022, 15, 5275

Figure 18. Flowchart of thermal analysis.

Figure 19. Temperature distribution.

Table 7. Material Properties for thermal analysis.

Part of the Machine Density (Kg/m3 ) Specific Heat (J/Kg◦ C) Thermal Conductivity (W/(m◦ C))
Modular Stator 7650 460 23
Mover 7650 460 23
Winding coils 4000 380 380

5.3. LPMEC Model


Software simulation of the thermal analysis is validated by an LPMEC model, taking
advantage of the heat and electrical system analogy, in which each part of the machine is
represented in the form of a resistor, and the value of the resistor depends on its areas and
specific heat capacity. A few assumptions were considered while developing this model,
and they are given below.

143
Energies 2022, 15, 5275

• Only two-dimensional heat flow is considered.


• All the material used in the machine are considered isotropic; heat travels equally in
all directions.
• Radiation of the heat is ignored.
The machine is divided into three major parts and each part is then represented by a
specific value resistance. Figure 20 represents the various parts of the machine. The coils
and magnets are divided into two exactly equal parts while the iron core is represented as
single part. Thermal conductivity of each part is presented in Table 8, while resistance of
each part can be calculated using Equations (10)–(14).

lix1 0.5 ∗ liy1


Ri1x = , Ri1y = (10)
λ f e liy1L a λ f e lix1L a

lix2 0.5 ∗ liy2


Ri2x = , Ri2y = (11)
λ f e liy2L a λ f e lix2L a
lix3 0.5 ∗ liy3
Ri3x = , Ri3y = (12)
λ f e liy3L a λ f e lix3L a
lix4 0.5 ∗ liy4
Ri4x = , Ri4y = (13)
λ f e liy4L a λ f e lix4L a
lmy
Rm = (14)
3λ PM lm x L a
Equivalent thermal resistance is considered in the case of the slot as the tempera-
ture distribution is not uniform because of the insulation and copper winding air gaps.
Equation (15) is used to find equivalent thermal resistance.

lc x 0.5lcy
Rcx = , Rcy = (15)
λcu lcy La λcu lc x La

Thermal capacity of the machine is represented in the form of capacitance in the


LPMEC model, with its values calculated from Equations (16) and (17).

C = m × cp (16)

m = ρ×V (17)

Figure 20. Division of the machine.

144
Energies 2022, 15, 5275

Table 8. Thermal conductivity of various parts.

Variable Thermal Conductivity


λfe 9 W/m◦ C
λcu 28 W/m◦ C
λ PM 0.68 W/m◦ C
λcu 28 W/m◦ C
Density of iron ρi 7870 Kg/m3
Density of copper ρc 8940 Kg/m3
Density of ferrite magnet ρ PM 5000 Kg/m3

The developed LPMEC model presented in Figure 21 is then simulated through


MATLAB Simulink software, and temperature rise of each part and the temperature rise of
the whole machine are observed, shown in Figure 22.

Figure 21. LPMEC model.

Figure 22. Thermal analysis of each part.

5.4. Comparison of FEA and LPMEC


Analysis of both FEA and LPMEC is compared, and a very good agreement between
the two is observed. Figure 22 shows thermal analysis of each part, according to the
division mentioned earlier, observed and compared with the average temperature rise of
FEA analysis. Figure 23 represents the comparison between the average temperature rise
from both methods. Overall, a relative difference of less than 4% is observed, which is quite

145
Energies 2022, 15, 5275

negligible, since the FEA analysis is performed in three dimensions and is therefore more
accurate, while in the case of LPMEC, we have only a two-dimensional flow of temperature.

Figure 23. Average temperature rise comparison.

6. Comparison with Conventional LHFSPM


Finally, the proposed LHFSPM is compared to the conventional design [26] in Table 9,
which includes a complete comparison. Both the machines use the same dimensions and
same type of PM, while the proposed machine achieves the same performance by reducing
the volume of PM by 25.56%.

Table 9. Comparison of proposed and conventional LHFSPM.

Parameter Proposed [26]


Mover length 131 mm
Stack length 90 mm
Airgap 0.8 mm
Rated velocity 4 m/s
DC current 5.76 A 6A
AC turns 230 276
DC turns 81 81
PM type Ferrite
PM volume 41.24 cm3 55.4 cm3
Thrust force 150.18 N 149.45 N

7. Conclusions
In this paper, a LHFSM with a crooked tooth modular stator is proposed. Electromag-
netic performance parameters such as no-load flux linkage, detent force, TF, and thrust
force profile at different velocities were studied and analyzed. The crooked tooth technique
was devised to improve the thrust force of the machine, and its effect at different angles
was presented. Geometric optimization technique was used to enhance the thrust force
and mitigate detent force and THD of the proposed machine. GA technique was used
to optimize the stator tooth, starting angle, and crooked angle. TF of the machine was
improved from an initial value of 129.53 N to an optimized value of 150.1823 N, TFd was
increased from 243.3 kN/m3 to 306.61 kN/m3 , ϕ( p− p) was increased from 0.12 T to 0.14 T,
Fd( p− p) was increased from 0.2601 N to 8.4 N, and the value for THD was slightly reduced
from the initial value. Iron losses for the machine were minimized significantly by the use
of the modular stator and optimal placement of the modules. In the end, thermal analysis
of the machine was performed and was then validated by the LPMEC model. A relative
percentage error of 3% was observed.

146
Energies 2022, 15, 5275

Author Contributions: Conceptualization, H.U.J. and F.K.; methodology, H.U.J.; software, H.U.J.,
B.U. and M.Q.; validation, H.U.J., B.U. and M.Q.; resources, F.K., A.H.M. and A.A.A.; original draft
preparation, H.U.J.; review and editing, B.U. and F.K.; supervision, F.K.; project administration,
F.K. and A.H.M.; funding acquisition, A.H.M. and A.A.A. All authors have read and agreed to the
published version of the manuscript.
Funding: There is no external funding.
Data Availability Statement: Not applicable.
Acknowledgments: This work was supported by COMSATS University Islamabad, Abbottabad
Campus.
Conflicts of Interest: The authors declare no conflicts of interest.

Abbreviations

ACw AC winding slot area


DCw DC winding slot area
DSLFSM Double-sided LFSM
G.A. Genetic algorithm
HEDSLFSM Hybrid excited double-sided linear flux switching machine
HELFSMs Hybrid excited LFSMs
HEVs Hybrid electric vehicles
LFSMs Linear flux switching machines
LIMs Linear induction machines
LPMC Lumped parametric magnetic equivalent circuit
PMLFSMs Permanent magnet LFSMs
S.R Split ratio
SLFSM Single-sided LFSM
STW Stator tooth width
TF Thrust force

References
1. Hu, D.; Xu, W.; Dian, R.; Liu, Y.; Zhu, J. Loss minimization control of linear induction motor drive for linear metros. IEEE Trans.
Ind. Electron. 2017, 65, 6870–6880. [CrossRef]
2. Lv, G.; Zeng, D.; Zhou, T.; Liu, Z. Investigation of forces and secondary losses in linear induction motor with the solid and
laminated back iron secondary for metro. IEEE Trans. Ind. Electron. 2016, 64, 4382–4390. [CrossRef]
3. Zou, J.; Xu, W.; Yu, X.; Liu, Y.; Ye, C. Multistep model predictive control with current and voltage constraints for linear induction
machine based urban transportation. IEEE Trans. Veh. Technol. 2017, 66, 10817–10829. [CrossRef]
4. Lv, G.; Zhou, T.; Zeng, D. Influence of the ladder-slit secondary on reducing the edge effect and transverse forces in the linear
induction motor. IEEE Trans. Ind. Electron. 2018, 65, 7516–7525. [CrossRef]
5. Sokolov, M.; Saarakkala, S.E.; Hosseinzadeh, R.; Hinkkanen, M. A dynamic model for bearingless flux-switching permanent-
magnet linear machines. IEEE Trans. Energy Convers. 2020, 35, 1218–1227. [CrossRef]
6. Cao, R.; Cheng, M.; Mi, C.C.; Hua, W. Influence of leading design parameters on the force performance of a complementary and
modular linear flux-switching permanent-magnet motor. IEEE Trans. Ind. Electron. 2013, 61, 2165–2175. [CrossRef]
7. Huang, W.; Hua, W.; Yin, F.; Yu, F.; Qi, J. Model predictive thrust force control of a linear flux-switching permanent magnet
machine with voltage vectors selection and synthesis. IEEE Trans. Ind. Electron. 2018, 66, 4956–4967. [CrossRef]
8. Huang, X.Z.; Yu, H.C.; Zhou, B.; Li, L.Y.; Gerada, D.; Gerada, C.; Qian, Z.Y. Detent-force minimization of double-sided permanent
magnet linear synchronous motor by shifting one of the primary components. IEEE Trans. Ind. Electron. 2019, 67, 180–191.
[CrossRef]
9. Hao, W.; Wang, Y. Thrust force ripple reduction of two c-core linear flux-switching permanent magnet machines of high thrust
force capability. Energies 2017, 10, 1608. [CrossRef]
10. Song, J.; Dong, F.; Zhao, J.; Lu, S.; Dou, S.; Wang, H. Optimal design of permanent magnet linear synchronous motors based on
taguchi method. IET Electr. Power Appl. 2017, 11, 41–48. [CrossRef]
11. Ma, W.; Wang, X.; Wang, Y.; Li, X. A hts-excitation modular fluxswitching linear machine with static seals. IEEE Access 2019, 7,
32009–32018. [CrossRef]
12. Wang, Z.-Q.; Long, Z.-Q.; Li, X.-L. Fault analysis and tolerant control for high speed pems maglev train end joint structure with
disturbance rejection. J. Electr. Eng. Technol. 2019, 14, 1357–1366. [CrossRef]

147
Energies 2022, 15, 5275

13. Mohamed Jaffar, M.Z. Modeling and Analysis of Saturated Inductances and Torque Ripple Components in Interior Permanent
Magnet Motors. Ph.D. Thesis, North Carolina State University, Raleigh, NC, USA, 2019.
14. Ullah, B.; Khan, F.; Milyani, A.H. Analysis of a Discrete Stator Hybrid Excited Flux Switching Linear Machine. IEEE Access 2022,
10, 8140–8150. [CrossRef]
15. Gas, P.; Kurgan, E. Evaluation of thermal damage of hepatic tissue during thermotherapy based on the arrhenius model. In
Proceedings of the 2018 Progress in Applied Electrical Engineering (PAEE), Koscielisko, Poland, 18–22 June 2018; pp. 1–4.
16. Zhao, J.; Guan, X.; Li, C.; Mou, Q.; Chen, Z. Comprehensive evaluation of inter-turn short circuit faults in pmsm used for electric
vehicles. IEEE Trans. Intell. Transp. Syst. 2020, 22, 611–621. [CrossRef]
17. Li, G.; Ojeda, J.; Hoang, E.; Gabsi, M.; Lecrivain, M. Thermal–electromagnetic analysis for driving cycles of embedded flux-
switching permanent-magnet motors. IEEE Trans. Veh. Technol. 2011, 61, 140–151. [CrossRef]
18. Thomas, A.; Zhu, Z.; Li, G. Thermal modelling of switched flux permanent magnet machines. In Proceedings of the 2014
International Conference on Electrical Machines (ICEM), Berlin, Germany, 2–5 September 2014; pp. 2212–2217.
19. Cai, X.; Cheng, M.; Zhu, S.; Zhang, J. Thermal modeling of fluxswitching permanent-magnet machines considering anisotropic
conductivity and thermal contact resistance. IEEE Trans. Ind. Electron. 2016, 63, 3355–3365. [CrossRef]
20. Jan, H.U.; Khan, F.; Ullah, B.; Qasim, M.; Khan, M.A.; Hafeez, G.; Albogamy, F.R. Design and Thermal Modeling of Modular
Hybrid Excited Double-Sided Linear Flux Switching Machine. Energies 2021, 14, 8511. [CrossRef]
21. Jan, H.U.; Khan, F.; Qasim, M.; Ullah, B.; Yousaf, M.; ul Islam, Z. Thermal and Stress Analysis of Linear Hybrid Excited Flux
Switching Machine with Modular Stator. In Proceedings of the 2021 International Conference on Emerging Power Technologies
(ICEPT), Topi, Pakistan, 10–11 April 2021; pp. 1–6. [CrossRef]
22. Ur Rahman, L.; Khan, F.; Khan, M.A.; Ahmad, N.; Khan, H.A.; Shahzad, M.; Ali, S.; Ali, H. Modular rotor single phase field
excited flux switching machine with non-overlapped windings. Energies 2019, 12, 1576. [CrossRef]
23. Liang, J.; Ming, Z.; Ji, X. Feeding apparatus directly driven by optimal topology flux switching permanent magnet linear motor.
Electr. Power Components Syst. 2019, 47, 903–913. [CrossRef]
24. Zeng, Z.; Shen, Y.; Lu, Q.; Wu, B.; Gerada, D.; Gerada, C. Investigation of a partitioned-primary hybrid-excited flux-switching
linear machine with dual-pm. IEEE Trans. Ind. Appl. 2019, 55, 3649–3659. [CrossRef]
25. Zhao, J.; Mou, Q.; Guo, K.; Liu, X.; Li, J.; Guo, Y. Reduction of the detent force in a flux-switching permanent magnet linear motor.
IEEE Trans. Energy Convers. 2019, 34, 1695–1705. [CrossRef]
26. Hwang, C.C.; Li, P.L.; Liu, C.T. Design and analysis of a novel hybrid excited linear flux switching permanent magnet motor.
IEEE Trans. Magn. 2012, 48, 2969–2972. [CrossRef]

148
energies
Article
Performance Analysis and Optimization of a Novel Outer Rotor
Field-Excited Flux-Switching Machine with Combined
Semi-Closed and Open Slots Stator
Siddique Akbar 1, *, Faisal Khan 1 , Wasiq Ullah 1 , Basharat Ullah 1 , Ahmad H. Milyani 2,3
and Abdullah Ahmed Azhari 4

1 Department of Electrical and Computer Engineering, COMSATS University Islamabad,


Abbottabad 22060, Pakistan
2 Department of Electrical and Computer Engineering, King Abdulaziz University, Jeddah 21589, Saudi Arabia
3 Center of Research Excellence in Renewable Energy and Power Systems, King Abdulaziz University,
Jeddah 21589, Saudi Arabia
4 The Applied College, King Abdulaziz University, Jeddah 21589, Saudi Arabia
* Correspondence: [email protected]

Abstract: Slotting effect in electric machines reduces flux per pole that effect magnetic flux density
distribution in the air gap which induces harmonics in magnetic flux density causing flux pulsation,
that in turn generates dominant torque pulsation in the form of cogging torque and torque ripples.
To overcome the abovesaid demerits, a novel outer rotor field-excited flux-switching machine (OR-
FSFSM) with a combined semi-closed and open slots stator is proposed in this study. The developed
OR-FEFSM offers a high-power factor, due to the utilization of the semi-closed slot for armature coils.
Citation: Akbar, S.; Khan, F.; Ullah, The open slot stator structure was chosen for the field excitation coil, which effectively suppresses
W.; Ullah, B.; Milyani, A.H.; Azhari, leakage reluctance that causes flux pulsation. Thus, the influence of torque ripples is reduced, and
A.A. Performance Analysis and the average torque is improved. In order to investigate the effectiveness of the proposed OR-FEFSM,
Optimization of a Novel Outer Rotor a detailed study of stator slot and rotor pole combinations are performed. Based on simplified
Field-Excited Flux-Switching mathematical formulation, 12S/7P (stator slot/rotor poles), 12S/11P, 12S/13P, and 12S/17P are the
Machine with Combined
most feasible combinations. Finite Element Analysis (FEA) based on comprehensive electromagnetic
Semi-Closed and Open Slots Stator.
performance is performed on each combination, and found that 12S/13P offers the highest average
Energies 2022, 15, 7531. https://
torque of 4.62 Nm, whereas 3.72 Nm, 2.72Nm, and 1.68 Nm average torque is offered by 12S/17P,
doi.org/10.3390/en15207531
12S/7P, and 12S/11P, respectively. Based on the initial analysis, 12S/13P was considered for further
Academic Editors: Vladimir Prakht, analysis and optimized using JMAG built-in Genetic Algorithm (GA). Moreover, thermal analysis
Mohamed N. Ibrahim and Vadim was performed, and the proposed design was compared with the conventional design.
Kazakbaev

Received: 16 September 2022 Keywords: field-excited machine; flux-switching machine; finite element analysis; genetic algorithm;
Accepted: 9 October 2022 thermal analysis; performance analysis; optimization
Published: 12 October 2022

Publisher’s Note: MDPI stays neutral


with regard to jurisdictional claims in
published maps and institutional affil-
1. Introduction
iations. Flux-switching machines (FSMs) are a class of synchronous machines that attract
researcher interest due to their doubly salient structure, featuring both excitation sources
(permanent magnet (PM) or winding) on the stator and robust rotor structure [1–5]. Since
there are no PMs or windings on the rotor, FSM has a robust rotor structure like switched
Copyright: © 2022 by the authors. reluctance machines (SRM), making it a strong candidate in many industrial applications
Licensee MDPI, Basel, Switzerland. where high speed and rigidness are required [6–8]. Thus, due to the inherent feature of
This article is an open access article high torque and power density, single-phase, three-phase, and multiphase FSM topologies,
distributed under the terms and
such as PM excited, field excited (FE), and hybrid excited (HE), have been researched over
conditions of the Creative Commons
the past few decades [9,10].
Attribution (CC BY) license (https://
Permanent magnet flux-switching machines (PMFSMs) offer high torque density, ef-
creativecommons.org/licenses/by/
ficiency, lower harmonic components in their back-electromotive force (back-EMF), and
4.0/).

Energies 2022, 15, 7531. https://doi.org/10.3390/en15207531 149 https://www.mdpi.com/journal/energies


Energies 2022, 15, 7531

robust rotor structure [11–13]. Due to the stator’s flux-concentrating effect, PMFSM eas-
ily generates a high magnetic flux density in the air gap and therefore achieves high
torque [14–16]. However, field weakening is a problem for these machines because their
d-axis inductance is lowered due to deep magnetic saturation in the stator, limiting their
constant power speed range. Additionally, the possibility of irreversible demagnetization,
cost of rare earth magnets, uncontrollable flux, and PM’s low mechanical strength are all
critical issues limiting PM machines to some extent. Therefore, reduced PM or no PM
machines are preferred [17,18].
Hybrid excited flux-switching machines (HEFSMs) have the potential to improve
power density, efficiency, and flux weakening performance [19–21], and overall reduce the
volume of PM by adding FE coils (FEC). However, the presence of both excitation sources
(PM and FEC) on the stator makes its geometry complex. To overcome the aforesaid
demerits of HEFSM and PMFSM, field-excited flux-switching machines (FEFSMs) are
introduced that are fully excited by windings and have more robust rotors than other types
of FSM. Despite their lower efficiency, the field current regulation makes FEFSM a suitable
alternative for variable speed applications [4,22].
FEFSM that utilize only field winding as a magnetic excitation are proposed in [23].
The field-weakening capabilities of this machine have been greatly enhanced due to flux
regulation capability with dc excitation current that effectively reduces excitation magnetic
field. However, the field windings are wound around every other stator tooth, causing the
magnetic field to be radially excited and unable to be concentrated, which results in low
torque density. To improve torque-power density, the authors of [23] investigated wound
field-excited FSM with non-overlapped winding and segmented rotor structure. Due to the
advantages of flexible flux weakening capability and short flux path, high-torque density
is achieved. However, FEFSM topologies with segmented rotors suffer from mechanical
constraints, and are not feasible for high-speed applications due to rotor segmentation.
The authors of [24] introduce a partitioned stator FEFSM (PS-FEFSM) with a double
stator, having field and armature winding on a separate stator, so that FEC are arranged in
the inner stator. In contrast, armature windings are placed in outer stators, thus increasing
the torque density by 19% more than conventional FEFSM. However, manufacturing cost
increases, due to the partitioned stator with two air gaps; and with metallic coupling of the
rotor pieces, the cage losses increase, which can be incorporated with non-metallic sticks.
Despite the average torque improvements, the influence of the cogging torque and
torque ripples are dominant in FEFSM. In this regard, the authors of [23] comparatively
investigated FEFSM with and without skew rotor for torque ripple reduction. However,
torque ripple is suppressed at the cost of a reduction in average torque. To further improve
the torque profile, the authors of [25] examined torque characteristics of an OR-FEFSM with
maximum torque control; however, this design suffers from magnetic saturation under
higher currents. Moreover, dual-stator OR-FEFSM for unbalanced shaft magnetic force
in [26] is effectively suppressed at 3% reduction in average torque.
To address the aforementioned problems in various state-of-the-art designs, this
paper proposes a novel OR-FEFSM (as shown in Figure 1) for direct-drive embedded
in-wheel systems, with an improved torque profile utilizing combined semi-closed and
open stators. Furthermore, due to the elimination of mechanical gearboxes, the proposed
OR-FEFSM omits maintenance costs, ultimately improving system operation efficiency.
In contrast, combined semi-closed and open slots dominantly suppress torque ripple
issues and improve overload capability. Therefore, this paper intends a finite element
analysis (FEA)-based detailed investigation of the proposed OR-FESM with possible rotor
pole combinations.

150
Energies 2022, 15, 7531

(a) (b)

Figure 1. Proposed OR-FEFSM (a) cross-sectional view and (b) magnetic flux density map.

Section 2 discusses design topology and operating principles, while Section 3 inves-
tigates FEA-based performance analysis of possible rotor pole combinations. Section 4
describes the genetic algorithm. Section 5 examines the best rotor pole’s detailed FEA-
based performance study. Section 6 depicts a quantitative comparison of electromagnetic
performance, whereas Section 7 investigates a comparison of conventional and proposed
designs. Section 8 draws a conclusion.

2. Design Topology and Operating Principles


2.1. Design Topology
The proposed 12S/13P OR-FEFSM is shown in Figure 1. It consists of a salient outer
rotor and an inner stator having both armature and field winding. The material used for
the stator and rotor is 35H210 silicon steel, and copper is used for the winding. Geometric
parameters of the proposed OR-FEFSM are shown in Figure 2 and listed in Table 1, while
the simplified operating principle based on the flux-switching principle and possible rotor
pole combinations are explored as follows.

Figure 2. Depicted design parameters of proposed OR-FEFSM.

2.2. Operating Principles


The operating principle of OR-FEFSM is similar to PMFSM, and can be easily explained
by the structure shown in Figure 3a,b. The upper laminated part is a rotor-like SRM, and
the lower part consists of an armature winding (AW) and FEC labeled in the stator. The
flux generated by the FEC flows through the path with the least reluctance. As shown in
Figure 3a, when the rotor poles align with the stator teeth, around which the phase coil is
wound, the flux that is linked in the coil leaves the rotor tooth and enters the coil. Similarly,
as the rotor rotates further and another rotor pole aligns with the next stator tooth of the
same coil, flux leaves the coil and enters the rotor tooth, maintaining the same amount of

151
Energies 2022, 15, 7531

flux-linkage but reversing the polarity, as shown in Figure 3b. The coils flux linkage varies
regularly as the rotor rotates, producing sinusoidal back-EMF. Torque is generated when
current is appropriately provided to the coils, which drives the armature forward.

Table 1. Design parameter of proposed OR-FEFSM.

Symbol Value (unit) Symbol Value (unit)


R1 15 mm Wry 7.5 mm
R2 20 mm Wrp 6 mm
R3 35.5 mm Hrp 7.5 mm
R4 40 mm rout 22 mm
R5 44.5 mm αin 10 mm
R6 45 mm Wsp 6 mm
R7 52.5 mm Hsi 15.5 mm
R8 60 mm Hso 9 mm
βin 6.76 mm β out 11.67 mm
W1 6 mm W2 6 mm
W3 6 mm Stack length 45 mm

(a) (b)

Figure 3. Operating principle of OR-FEFSM (a) positive maximum flux and (b) negative maxi-
mum flux.

2.3. Rotor Poles Combination


Rotor pole position determines the basic principle of FEFSM due to magnetic flux link-
age in armature winding that will be positive and negative, depending on the rotor position.
For inner and outer rotor FEFSM, possible rotor poles combinations are computed as:
 
1±k
Nr= Ns (1)
2q

where Nr represents the number of rotor poles, Ns stator slot, K is natural number, and q is
number of pairs of phases. For the proposed OR-FEFSM, the values of q and Ns are 6 and
12, respectively, whereas feasible stator slot and rotor pole obtained are 12S/7P, 12S/11P,
12S/13P, and 12S/17P.
Furthermore, for all possible rotor poles combinations, the number of turns per phase
(Na ) and number of turns for field winding (Ne ) varies based on the armature current
density (Ja ) and field excitation current density (Je ) as follows:

Ia Na
Ja = (2)
αSa
Ie Ne
Je= (3)
αSe

152
Energies 2022, 15, 7531

where α represent the filling factor while Sa and Se represent armature and field slot area,
respectively. Thus, to decide an optimal values for the aforesaid associated parameters
(Je & Ja ), a detailed electromagnetic performance is investigated.

3. Performance Analysis of Feasible Rotor Pole Combination


In this section, a comparative analysis of the proposed OR-FEFSM is conducted for
the possible rotor pole combinations. The best design based on electromagnetic analysis
will be proceeded for further study.

3.1. No Load Flux Linkage


Figure 4 shows the no-load flux linkage of all feasible rotor pole combinations.
It is worth noting that the phase flux linkage obtained under no-load operation with
Je = 15 A/mm2 . Analysis reveals that OR-FEFSM with 12S-7P design offers the highest
flux linkage of 0.020 Wb, which is 54%, 7.15%, and 37.97% higher than 11, 13, and 17 rotor
poles, respectively.
)OX[OLQNDJH :E

Figure 4. No-load operation flux linkages of all feasible combinations.

3.2. No Load Back-EMF


Under no-load operation, phase back-EMF for 12S/7P, 12S/11P, 12S/13P, and 12S/17P
through 2D FEA is shown in Figure 5. Analysis reveals that 12S/13P shows the highest
back-EMF of 44.43 V, whereas the phase back-EMF for 12S/7P, 12S/11P, and 12S/17P are
25 V, 18.13 V, and 38.63 V, respectively.
%DFN(0) 9

Figure 5. No-load operation phase back-EMF.

153
Energies 2022, 15, 7531

3.3. Cogging Torque


The torque obtained is cogging torque when no current is applied to the armature
winding. This torque is due to slotting effects that cause acoustic noise and vibration, and
is therefore undesirable. Since slotting effects change with the rotor pole number variation,
the resultant cogging torque greatly vary. The cogging torques of 12S/7P, 12S/11P, 12S/13P,
and 12S/17P are illustrated in Figure 6. In regard to peak magnitude, analysis reveals that
12S/17P design offers lowest cogging torque of 0.0508 Nm, whereas 12S/11P shows highest
cogging torque of 0.337 Nm. Furthermore, the cogging torque profiles for 12S/13P and
12S/7P show moderate values.

Figure 6. Variation of cogging torque profile.

3.4. Instantenous Torque


Under loaded and rated operating conditions of Ja = Je = 15 A/mm2 and speed
of 1500 rpm, instantaneous torque profile is shown in Figure 7. Comparative analysis 
of various rotor poles illustrates that 12S/13P offers the highest average torque Tavg of
4.62 Nm, whereas 12S/11P shows the lowest average torque of 1.68 Nm.
,QVWDQWDQHRXVWRUTXH 1P

Figure 7. Instantaneous torque profile with rotor pole number.

Table 2 represents average torque, cogging torque, torque ripples ratio, flux linkage,
and back-EMF of possible rotor pole combinations. From the above analysis, it is clear
that the 12S/13P combination has a more significant torque density and back-EMF at
Ja = Je = 15 A/mm2 . Therefore, it is considered for further study.

154
Energies 2022, 15, 7531

Table 2. Quantitative performance of the OR-FEFSM with various rotor pole numbers at
Ja = Je = 15 A/mm2 .

12S/7P 12S/11P 12S/13P 12S/17P


Average Torque (Nm) 2.72 1.68 4.62 3.72
Cogging Torque P-P (Nm) 0.344 0.377 0.246 0.0508
Torque ripples (%) 32 16 31 12
Flux Linkage (Wb) 0.022816 0.010481 0.021185 0.014152
Back-EMF (V) 25.06 18.13 44.84 38.64

4. Genetic Optimization
In this section, optimization using genetic algorithm (GA) is used for the proposed
OR-FEFSM with 12S/13P. GA is a widely common solution for optimization issues because
it is particularly good at solving highly non-linear objective functions. The GA-based
optimization technique is based on the evolution of the world’s animal population. It begins
with a random population of variables, similar to a pool of chromosomes. A generation
is the term used to describe each iteration of the algorithm. Only a few chromosomes
with the most significant fitness values are carried down into the next generation, known
as exceptional offspring. In addition, the technique generates new children for the next
generation by simulating crossover and mutation, which are binary and unary acts on
existing chromosomes, respectively. The method is repeated until one of the end conditions
are met [27]. The number of generations and population size are defined to reach an
optimum global value of the objective function. Figure 8 illustrates the GA workflow.

67$57

'HILQHLQLWLDOSDUDPHWHUV

,PSRUW&$'PRGHO

'HILQHREMHFWLYHIXQFWLRQ

6HWFRQVWUDLQV

6SHFLI\SRSXODWLRQVL]HDQG
&URVVRYHU
QXPEHURIJHQHUDWLRQ

1XPEHURI ([HFXWH)($
FKLOGUHQV

&KHFN 1R 2EMHFWLYHIXQFWLRQ
FRQVWUDLQV FRQYHUJHV"

<HV

1R 1H[W
*OREDOO\RSWLPL]HG"
JHQHUDWLRQ

<HV

6723

Figure 8. Flow chart of Genetic Optimization.

Geometry editor is used to develop the initial design, and the CAD parameters are
then imported into the designer. The objective function and constraints are given below,

155
Energies 2022, 15, 7531

composed of two sub-objectives. Table 3 describes the ranges and constraints of variable
CAD parameters.

max( Tavg )
Objective function =
min( Tripples )

Tavg > 4.62
Constraints =
Tripples < 0.28


⎪ 3 ≤ Wry ≤ 12


⎪ 2 ≤ Wrp ≤ 5


3 ≤ W1 ≤ 8
Design variables =

⎪ 3 ≤ W2 ≤ 8



⎪ 3 ≤ W3 ≤ 8

12 ≤ R1 ≤ 15

Table 3. Variables for Genetic Optimization.

Parameters Initial (mm) Boundary Conditions Optimized (mm)


Wry 7.5 3 ≤ Wry ≤ 12 6.3505
Wrp 3 1 ≤ Wrp ≤ 5 3.3055
W1 6 3 ≤ W1 ≤ 8 7.97017
W2 6 3 ≤ W2 ≤ 8 5.9480
W3 6 3 ≤ W3 ≤ 8 6.8116

JMAG built-in global optimization, which utilizes the GA approach, was used to
optimize the geometrically significant parameters. Geometrical design variables, such as
the yoke length (Wry ), width of rotor tooth (Wrp ), the area of the armature slot, area of
the FEC slot, width of stator poles (W 1 , W 2 , W 3 ), and the shaft radius (R1 ), were used to
determine the optimization problem of the design. During optimization, key dimensions,
such as the stator outer radius (R5 ), air gap, rotor outer radius (R8 ), stack length, rated
field and armature current densities, and turns, were kept constant to maintain constant
electrical and magnetic loading. In optimization settings, maximum generations were set
at 14, population size at 16, number of children at 17, and stopping criteria at 10. The
number of elements and nodes of each model were 12,246 and 7672, respectively. The mesh
size was set to 0.5. After computing 451 case studies with GA, which took nearly 112 hours,
the file size was 110 GB, the PC used was a HP core i5, 2.5 GHz, 8 GB RAM, and optimum
values were obtained. Optimization results, such as torque and ripple ratio w.r.t, Wry , Wrp ,
W 1 , W 2 , W 3 , and R1 , are illustrated in Figure 9 in relation to one another. Table 3 depicts
the proposed design’s global optimal parameters.

Figure 9. Cont.

156
Energies 2022, 15, 7531

7RUTXHULSSOHUDWLR 
7RUTXHULSSOHUDWLR 
$YHUDJHWRUTXH 1P
$YHUDJHWRUTXH 1P

Figure 9. Torque and ripples ratio w.r.t optimization parameters (a) R1 (b) Wry (c) W1 (d) W2 (e) W3
(f) Wr .

5. FEA Based Electromagnetic Performance of Optimized Design


In this section, a detailed comparative analysis of optimal design is discussed.

5.1. No Load Analysis


The no-load airgap flux density of 12S/13P initial and optimized design substantially
fluctuates, resulting in a significant variation in the magnetic flux density over one periodic
cycle, as shown in Figure 10. The peak value of airgap flux density of optimized design is
30% more than the initial. The variation of the magnetic flux density results in phase flux
linkage variation. Figure 11a,b shows no load flux linkage and harmonic spectra of initial
and optimized design. Analysis shows that the peak value of flux linkage is improved
by 31.15% after optimization. Similarly, odd harmonics and total harmonic distortion
(THD) were improved by 41.88 and 58.33%. The formula for calculating THD is shown in
(4) below.
∑ N ∅i
THD = i=2 (4)
∅1
where ∅1 is fundamental hormonic component and N is natural number

157
Energies 2022, 15, 7531

Figure 10. Variation of the magnetic flux density over one periodic cycle.
3KDVH)OX[/LQNDJH :E

Figure 11. Initial and optimized design phase flux linkage (a) waveform and (b) harmonic spectra.

Under no-load condition, phase-back-EMF and harmonic spectra for initial and opti-
mized design through 2D FEA is shown in Figure 12a,b. Analysis reveals that back-EMF is
improved by 28%, while odd harmonics and THD is improved by 33.2% and 39.63%, respec-
tively, after optimization. Figure 13 shows the comparison between initial and optimized
cogging torque; after optimization, the cogging torque is reduced by 54.93%.


,QLWLDOGHVLJQ
 2SWLPL]HGGHVLJQ










           
+DUPRQLFV2UGHU
E

Figure 12. Initial and optimized design back-EMF (a) waveform and (b) harmonic spectra.

158
Energies 2022, 15, 7531

Figure 13. Cogging torque analysis of initial and optimized design.

5.2. Load Analysis


Under loaded and rated operating conditions of Ja = Je = 15 A/mm2 and speed of
1500 rpm, instantaneous torque profile of 12S/13P is shown in Figure 14, which offers
a 41.99% improvement in torque and 87.76% in Torque ripple ratio ( Trr ). Furthermore,
average torque profiles with torque ripples under different electric loading are illustrated
in Figure 15, demonstrating the over-load capability of the proposed OR-FEFSM before
and after optimization. It can be seen that under all-electric loading conditions, average
torque increases, whereas the torque ripple ratio decreases. From the analysis, the torque
and ripples ratios are improved by 41.99% and 87.76%, respectively, at the rated conditions.
The formula for calculating torque ripple ratio is discussed in (5).

Tmax − Tmin
Trr = × 100 (5)
Tavg

Figure 14. Instantaneous toque profile of initial and optimized design.

159
Energies 2022, 15, 7531

7RUTXHULSSOHUDWLR 
$YHUDJHWRUTXH 1P

7RUTXHULSSOHUDWLR 
$YHUDJHWRUTXH 1P

Figure 15. Overload capability of proposed design before and after optimization. (a) Torque w.r.t Ja
and Je before optimization; (b) ripples w.r.t Ja and Je before optimization; (c) torque w.r.t Ja and Je
after optimization; and (d) ripples w.r.t Ja and Je after optimization.

In (5) Tmax is the maximum value of torque, Tmin is the minimum value of torque,
and Tavg is the average value of torque. Figure 16 shows electromagnetic torque perfor-
mance of initial and optimized design w.r.t different electrical degrees.
$YHUDJHWRUTXH 1P

Figure 16. Torque w.r.t different electrical degrees before and after optimization.

160
Energies 2022, 15, 7531

5.3. Dynamic Analysis


In designing an electric machine, dynamic characteristics analysis is one of the key
studies that provides wide operational capability. In this regard, torque and output power,
characteristics curve versus speed and power factor, give a detailed illustration of its
behavior under low-speed, high-speed, constant torque, and constant output power regions,
which is a prerequisite for direct-drive in-wheel-embedded systems.
The output power of the proposed machine is calculated by multiplying Tavg with
the corresponding speed. Input power is the summation of output power and total losses,
including core and copper losses. Core losses can be calculated from 2D-FEA at the specified
points, and (6) is used for copper losses calculation.

PCopper = PCopper ( AW ) + PCopper ( FEC ) (6)

where,
PCopper ( AW ) = Irms ρJL( NQ)(1000) (7)
PCopper ( FEC) = IρJL( NQ)(1000) (8)
In (8), I represent current, ρ is resistivity, J represents current density, N and Q are
number of turns and slot pair, respectively.

Pout = Tavg ∗ ω (9)

Pinput = Pout + PCopper + PCore (10)


Finally, efficiency is given as:

Pout
η= 100 % (11)
Pinput

The characteristics curve of torque and power versus speed on initial and optimized
designs is depicted in Figures 17a and 17b, respectively.
$YHUDJHWRUTXH 1P

2XWSXWSRZHU N:

Figure 17. (a) Torque vs. speed curve; (b) power vs. speed curve.

Similarly, the power factor can be calculated as:

3 
P= V Iq + Vq Id (12)
2 d
3 
Q= V Iq + Vq Id (13)
2 d

S = P2 + Q2 (14)

161
Energies 2022, 15, 7531

P
PF = cos∅ = (15)
S
The power factor of initial and optimized design is shown in Figure 18.

Figure 18. Power factor of initial and optimized design w.r.t different speed.

A comparison of the initial and optimized OR-FEFSM designs is performed in order to


verify and examine the performance of the GA optimization technique. Table 4 illustrates a
quantitative assessment of the proposed machine.

Table 4. Performance comparison of initial and optimized design.

Parameters Initial Value Optimized Value Improvement


Torque (Nm) 4.877 6.925 41.99%
Torque Ripples (%) 0.30 0.037 87.76%
Flux linkage (Wb) 0.021 0.0276 23.91%
Airgap flux density (T) 1.3886 1.9276 27.94%
Back − emf peak (V) 41.07 55.362 25.91%
Cogging torque(Nm) 0.3160 0.1424 54.93%
Odd harmonics in flux (Wb) 0.000573 0.000333 41.88%
THD in flux (%) 2.88 1.2 58.33%
Odd harmonics in back-EMF (V) 6.1206 4.09209 33.2%
THD in back emf (%) 16.4 9.9 39.63%
Output power (Kw) 2.27 3.223 29.56%
Power factor (Pf) 0.84 0.97 13.4%
Efficiency (%) 83.4 87.70 4.9%

6. Thermal Analysis
Thermal analysis is one of the most important analyses in electric machine design, as
heat dissipation limits the long-term functionality of the machine. Therefore, it helps to
determine the operating limits and the insulating class of the machine. The electromagnetic
performance reduces when the temperature exceeds a certain permissible range, resulting
in an inter-turn short circuit problem [28]. Current flow through the conductor produces
heat dissipation and power loss, which raises the temperature of the whole machine. The
machine has core losses in addition to the copper losses associated with the windings, which
are due to the eddy current and hysteresis losses in the core. All of these losses behave as a
heat source and raise temperatures. Therefore, the magnetic loss study is used to determine
the losses, and thermal analysis is used to analyze the temperature distribution.
For the proposed machine, initial losses were determined using a 3D FEA followed
by thermal analysis in 3D thermal studies, as 3D analysis is more precise than 2D analy-
sis. To calculate the temperature distribution of the complete machine, the 3D loss study

162
Energies 2022, 15, 7531

is combined with the 3D heat study. The 3D thermal study reveals that the stator tem-
perature significantly rises due to the presence of all excitation sources on the stator;
however, the temperature of the rotor minimally rises, as the proposed machine has a
salient rotor. Figure 19 depicts a contour plot of the temperature distribution. In the stator,
the temperature distribution in the windings reaches a maximum value of 46 ◦ C. At the
rotor, the highest temperature observed is 24 ◦ C degrees.

Figure 19. Thermal analysis of the proposed OR-FEFSM.

7. Comparison of Proposed and Conventional Design


In this section, a detailed comparison of the conventional and proposed design is
performed. For fair comparison with [23], the proposed model is redesigned and optimized
under the same outer dimensions, stack length, current density, and air gap length. The
average torque of scaled 12S/13P design is 47.95 Nm which is 36.6%, 57.1%, 48.77%,
and 96.52% more than the conventional straight WFSM, skewed WFSM, 12S/8P SRM,
and 12S/10P SRM, respectively. Based on the key performance matrix listed in Table 5,
without any influence on efficiency, it is clear that the proposed design offers 9.4%, 65.8%,
and 63.7% less ripples than the conventional straight WFSM, 12S/8P SRM, and 12S/10P
SRM, respectively. Furthermore, individual machine part weight and overall weight
of the proposed machine are the lowest, resulting in a better torque density than the
aforementioned state-of-the-art model.
Table 5. Performance comparison of conventional and proposed design.

Straight Skewed 12/8 12/10 Proposed


Parameters
WFSM WFSM SRM SRM Machine
Average Torque (Nm ) 35.1 30.52 32.32 28.4 47.95
Torque Ripple (%) 27.6 6.85 73.1 68.9 25
Efficiency (%) 89.4 88 89.9 90.5 89.3
Weight of iron core (kg) 19.87 19.54 19.74 16.19
Weight of copper (kg) 17.86 13.74 12.63 3.121
Weight in total (kg) 37.73 34.33 32.37 19.311
Torque Density (Nm/kg ) 0.930 0.808 0.941 0.877 2.483

8. Conclusions
This paper proposed a novel outer rotor field-excited flux-switching machine with a
combined semi-closed and open slots stator to improve power factor and torque profile.

163
Energies 2022, 15, 7531

Semi-closed slots utilization improves power factor, whereas open slots suppress slot
leakages that effectively suppress leakage reluctance that causes flux pulsation, improving
torque ripples, and mean torque. The effectiveness of the proposed OR- FEFSM was
investigated through FEA with different stator slot and rotor pole combinations. Initial
FEA-based comprehensive electromagnetic performance reveals that 12S/13P offers the
highest average torque of 4.62 Nm, which was considered for further study and optimized
through GA optimization. The maximum output power obtained is 3.22 kW with an
efficiency of 87.7%. Finally, the proposed design was compared with the state-of-the-art
model, and the results showed there was significant improvement in average torque, torque
ripples, weight of machine, and torque density.

Author Contributions: Conceptualization, S.A., W.U. and F.K.; methodology, S.A. and B.U.; software,
S.A. and B.U.; validation, S.A. and W.U.; resources, F.K., A.H.M. and A.A.A.; formal analysis, B.U.,
A.H.M. and A.A.A.; original draft preparation, S.A. and W.U.; visualization, B.U.; review and editing,
B.U. and F.K.; supervision, F.K.; project administration, F.K., A.H.M. and A.A.A.; funding acquisition,
A.H.M., A.A.A. All authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Ullah, W.; Khan, F.; Sulaiman, E.; Umair, M.; Ullah, N.; Khan, B. Analytical validation of novel consequent pole E-core stator
permanent magnet flux switching machine. IET Electr. Power Appl. 2020, 14, 789–796. [CrossRef]
2. Jiang, W.; Huang, W.; Lin, X.; Zhao, Y.; Zhu, S. Analysis of Rotor Poles and Armature Winding Configurations Combinations of
Wound Field Flux Switching Machines. IEEE Trans. Ind. Electron. 2021, 68, 7838–7849. [CrossRef]
3. Jiang, W.; Huang, W.; Lin, X.; Zhao, Y.; Wu, X.; Zhao, Y.; Dong, D.; Jiang, X. A Novel Stator Wound Field Flux Switching Machine
with the Combination of Overlapping Armature Winding and Asymmetric Stator Poles. IEEE Trans. Ind. Electron. 2022, 69,
2737–2748. [CrossRef]
4. Zulu, A.; Mecrow, B.; Armstrong, M. A Wound-Field Three-Phase Flux-Switching Synchronous Motor with All Excitation Sources
on the Stator. IEEE Trans. Ind. Appl. 2010, 46, 2363–2371. [CrossRef]
5. Tang, Y.; Paulides, J.; Motoasca, T.; Lomonova, E. Flux-Switching Machine with DC Excitation. IEEE Trans. Magn. 2012, 48,
3583–3586. [CrossRef]
6. Wu, Z.; Zhu, Z.; Wang, C. Reduction of On-Load DC Winding-Induced Voltage in Partitioned Stator Wound Field Switched Flux
Machines by Dual Three-Phase Armature Winding. IEEE Trans. Ind. Electron. 2022, 69, 5409–5420. [CrossRef]
7. Ullah, W.; Khan, F.; Hussain, S. A Comparative Study of Dual Stator with Novel Dual Rotor Permanent Magnet Flux Switching
Generator for Counter Rotating Wind Turbine Applications. IEEE Access 2022, 10, 8243–8261. [CrossRef]
8. Tang, Y.; Motoasca, E.; Paulides, J.; Lomonova, E. Comparison of flux-switching machines and permanent magnet synchronous
machines in an in-wheel traction application. COMPEL-Int. J. Comput. Math. Electr. Electron. Eng. 2012, 32, 153–165. [CrossRef]
9. Othman, S.; Ahmad, M.; Rahim, J.; Bahrim, F.; Sulaiman, E. Design Improvement of Three Phase 12Slot-14Pole Outer Rotor Field
Excitation Flux Switching Motor. Int. J. Power Electron. Drive Syst. (IJPEDS) 2017, 8, 239. [CrossRef]
10. Shen, Y.; Hu, P.; Jin, S.; Wei, Y.; Lan, R.; Zhuang, S.; Zhu, H.; Cheng, S.; Chen, J.; Wang, D.; et al. Design of Novel Shaftless
Pump-Jet Propulsor for Multi-Purpose Long-Range and High-Speed Autonomous Underwater Vehicle. IEEE Trans. Magn. 2016,
52, 1–4. [CrossRef]
11. Su, P.; Hua, W.; Wu, Z.; Chen, Z.; Zhang, G.; Cheng, M. Comprehensive Comparison of Rotor Permanent Magnet and Stator
Permanent Magnet Flux-Switching Machines. IEEE Trans. Ind. Electron. 2019, 66, 5862–5871. [CrossRef]
12. Hua, W.; Cheng, M.; Zhu, Z.; Howe, D. Analysis and Optimization of Back EMF Waveform of a Flux-Switching Permanent
Magnet Motor. IEEE Trans. Energy Convers. 2008, 23, 727–733. [CrossRef]
13. Du, Y.; Zhang, C.; Zhu, X.; Xiao, F.; Sun, Y.; Zuo, Y.; Quan, L. Principle and analysis of doubly salient PM motor with Π-shaped
stator iron core segments. IEEE Trans. Ind. Electron. 2018, 66, 1962–1972. [CrossRef]
14. Ullah, W.; Khan, F.; Hussain, S. A Novel Dual Rotor Permanent Magnet Flux Switching Generator for Counter Rotating Wind
Turbine Applications. IEEE Access 2022, 10, 16456–16467. [CrossRef]
15. Chau, K.; Chan, C.C.; Liu, C. Overview of permanent-magnet brushless drives for electric and hybrid electric vehicles. IEEE Trans.
Ind. Electron. 2008, 55, 2246–2257. [CrossRef]
16. Du, Z.S.; Lipo, T.A. Efficient utilization of rare earth permanent-magnet materials and torque ripple reduction in interior
permanent-magnet machines. IEEE Trans. Ind. Appl. 2017, 53, 3485–3495. [CrossRef]
17. Dorrell, D.; Parsa, L.; Boldea, I. Automotive Electric Motors, Generators, and Actuator Drive Systems with Reduced or No
Permanent Magnets and Innovative Design Concepts. IEEE Trans. Ind. Electron. 2014, 61, 5693–5695. [CrossRef]

164
Energies 2022, 15, 7531

18. Boldea, I.; Tutelea, L.N.; Parsa, L.; Dorrell, D. Automotive Electric Propulsion Systems with Reduced or No Permanent Magnets:
An Overview. IEEE Trans. Ind. Electron. 2014, 61, 5696–5711. [CrossRef]
19. Yang, G.; Lin, M.; Li, N.; Tan, G.; Zhang, B. Flux-Weakening Control Combined with Magnetization State Manipulation of Hybrid
Permanent Magnet Axial Field Flux-Switching Memory Machine. IEEE Trans. Energy Convers. 2018, 33, 2210–2219. [CrossRef]
20. Wardach, M.; Palka, R.; Paplicki, P.; Prajzendanc, P.; Zarebski, T. Modern Hybrid Excited Electric Machines. Energies 2020, 13, 5910.
[CrossRef]
21. Zhang, G.; Hua, W.; Cheng, M.; Liao, J.; Wang, K.; Zhang, J. Investigation of an improved hybrid-excitation flux-switching
brushless machine for HEV/EV applications. IEEE Trans. Ind. Appl. 2015, 51, 3791–3799. [CrossRef]
22. Pollock, C.; Pollock, H.; Barron, R.; Coles, J.R.; Moule, D.; Court, A.; Sutton, R. Flux-switching motors for automotive applications.
IEEE Trans. Ind. Appl. 2006, 42, 1177–1184. [CrossRef]
23. Zhao, G.; Hua, W.; Qi, J. Comparative Study of Wound-Field Flux-Switching Machines and Switched Reluctance Machines.
IEEE Trans. Ind. Appl. 2019, 55, 2581–2591. [CrossRef]
24. Zhu, Z.Q.; Wu, Z.Z.; Evans, D.J.; Chu, W.Q. A Wound Field Switched Flux Machine with Field and Armature Windings Separately
Wound in Double Stators. IEEE Trans. Energy Convers. 2015, 30, 772–783. [CrossRef]
25. Yang, S.; Zhang, J.; Jiang, J. Modeling Torque Characteristics and Maximum Torque Control of a Three-Phase, DC-Excited
Flux-Switching Machine. IEEE Trans. Magn. 2016, 52, 1–4. [CrossRef]
26. Nguyen, H.Q.; Jiang, J.Y.; Yang, S.M. Design of a wound-field flux switching machine with dual-stator to reduce unbalanced shaft
magnetic force. J. Chin. Inst. Eng. 2017, 40, 441–448. [CrossRef]
27. Parappagoudar, M.B.; Pratihar, D.K.; Datta, G.L. Forward and reverse mappings in green sand mould system using neural
networks. Appl. Soft Comput. 2008, 8, 239–260. [CrossRef]
28. Zhao, J.; Guan, X.; Li, C.; Mou, Q.; Chen, Z. Comprehensive evaluation of inter-turn short circuit faults in pmsm used for electric
vehicles. IEEE Trans. Intell. Transp. Syst. 2020, 22, 611–621. [CrossRef]

165
energies
Article
Hybrid Switched Reluctance Motors for Electric Vehicle
Applications with High Torque Capability without
Permanent Magnet
Vijina Abhijith 1, *, M. J. Hossain 1, *, Gang Lei 1 , Premlal Ajikumar Sreelekha 2, *,
Tibinmon Pulimoottil Monichan 2 and Sree Venkateswara Rao 3

1 School of Electrical and Data Engineering, University of Technology Sydney, Ultimo, NSW 2007, Australia
2 Entuple E-Mobility, Diamond District, HAL Old Airport Road, Kodihalli, Bangaluru 560008, India
3 Gyrate the Motor Inside, Shimla Block, Hill country, Bachupallay, Telangana 500090, India
* Correspondence: [email protected] (V.A.); [email protected] (M.J.H.);
[email protected] (P.A.S.)

Abstract: In electric vehicle (EV) applications, hybrid excitation of switched reluctance motors
(HESRMs) are gaining popularity due to their advantages over other EV motors. The benefits include
control flexibility, simple construction, high torque/power density, and the ability to operate over a
broad speed range. However, modern HESRMs are constructed by increasing the air gap flux density
with permanent magnets (PMs) in the excitation system in order to generate more electromagnetic
torque. This study aims to investigate a new topology for increasing the torque capabilities of HESRM
without the use of permanent magnets (PMs) or other rare-earth components. This paper provides
a comprehensive evaluation of the static and dynamic characteristics, software analysis using the
Citation: Abhijith, V.; Hossain, M.J.; Ansys 2D finite element method (FEM), and an experimental demonstration of the real-time motor
Lei, G.; Sreelekha, P.A.; Monichan,
with an advanced control strategy in MATLAB/Simulink. Our simulation and experimental results
T.P.; Rao, S.V. Hybrid Switched
for a machine with 12/8 poles and a machine rating of 1.2 kW indicate that the HESRM designed
Reluctance Motors for Electric
without PMs has greater torque capability and efficiency than the conventional SRM. The proposed
Vehicle Applications with High
Torque Capability without
HESRM without PMs has a high torque/power density and a higher torque per ampere across the
Permanent Magnet. Energies 2022, 15, entire speed range, making it suitable for EV applications.
7931. https://doi.org/10.3390/
en15217931 Keywords: switched reluctance motor (SRMs); hybrid excitation of SRM (HESRM); hybrid excitation
of SRM without a permanent magnet; permanent magnet (PM); electric vehicle motor
Academic Editors: Vladimir Prakht,
Mohamed N. Ibrahim and Vadim
Kazakbaev

Received: 24 September 2022 1. Introduction


Accepted: 21 October 2022
The widespread use of ICE (internal combustion engine) automobiles has led to an
Published: 26 October 2022
increase in pollution and energy constraints in recent years. The use of electric vehicles
Publisher’s Note: MDPI stays neutral (EVs) has the potential to reduce environmental stress, reduce pollution, and limit the
with regard to jurisdictional claims in number of vehicles requiring charging [1]. Electric vehicles (EVs) rely heavily on electric
published maps and institutional affil- motors [2]; thus, numerous electric motors were evaluated and analyzed to determine
iations. which would be the most effective for various applications. These days, brushless DC
motors (BLDCMs), induction motors, and different permanent magnet (PM) machines are
some of the motor types other than SRM motors that can be used effectively in variable-
speed applications [3,4]. In comparison with conventional motors, SRMs are appealing
Copyright: © 2022 by the authors.
to researchers due to their low complexity, high durability, cost-effectiveness, and ease
Licensee MDPI, Basel, Switzerland.
of production [5,6]. SRM appears to have the same power density as induction motors,
This article is an open access article
and its reliability appears to be far superior to that of DC series and permanent magnet
distributed under the terms and
motors [7]. As the SRM lacks the mechanical components of other conventional motors,
conditions of the Creative Commons
such as permanent magnets and rotor windings, it can be utilized in high-temperature
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
environments [8]. Low torque and power density, compared with permanent magnet
4.0/).
motors, as well as significant torque ripple, are two drawbacks that limit its utility [9].

Energies 2022, 15, 7931. https://doi.org/10.3390/en15217931 167 https://www.mdpi.com/journal/energies


Energies 2022, 15, 7931

Numerous researchers are devoted to reducing the machine architecture to the level
required by applications in order to address these problems and boost the performance
of SRMs. Prior research has demonstrated that numerous approaches to switched reluc-
tance motors focus on innovative controller circuits [10–12], core segmentation [13–16],
and hybrid excitations (HESRMs).
Using sophisticated controller circuits, the switching circuit of the stator of an SRM
can be optimized to provide maximum torque. Precision phase-winding switching makes
it possible to tailor the performance of modern electronic converters to specific applica-
tions [17]. This accurate switching method permits substantial control over the torque
ripple of the motor. In order to increase the available flux in the air gap and the resulting
flux, the segmentation method [18] is applied to either the stator or rotor cores. By limiting
the flux to the separated regions, the average torque of the active region is increased. This
segmented system diminishes mechanical strength and contributes to higher costs [19]
due to the increased production complexity. Due to the internal flux’s reciprocal coupling,
the overall efficiency of the segmented core is diminished. Drive circuits with multiple
stator segments may be used for low- and high-speed applications and provide more
power and torque per ampere [20]. This technique increases torque output at the expense
of machine durability [21]. The third option (HESRM) is permanent magnet hybrid-excited
switched reluctance motors.
SRMs underwent a paradigm shift with the introduction of the hybrid excitation
technique, which substantially increased their ability to generate electromagnetic torque.
However, this approach uses PMs into the structure of the machine to increase the motor’s
average torque. A 12/8 HESRM with a modular stator was presented in a stator segmented
with PMs [22]. There is evidence that it can transport a maximum flux density at a higher
torque [23]. The flux capacity is increased by 34.7% compared with the segmented SRM
without PM [24]. As a result, the average torque is 71.4% greater than the conventional
SRM. A comparison study of a segmented stator permanent magnet hybrid excitation
SRM drive [25] reported that the static average torque of the segmented HESRM is 1.16
to 1.39 times that of the standard one. There were six permanent magnets (PMs) in the
stator of the machine; due to the behavior of the PMs, this could cause cogging torque.
It has been demonstrated that SRM with auxiliary winding and permanent magnets is
3% more efficient and reduces size and weight by the same amount [26]. The HESRM,
meanwhile, combines the benefits of the permanent magnet machine and the electromagnet
machine to create a hybrid excitation of the double salient machine [27]. The rare-earth
permanent magnets in the rotor core of the 6-slot 8-pole hybrid excitation flux switching
machine (HEFSM) are an innovative design element that contributes to the machine’s high
power density by providing adequate mechanical strength with enhanced electromagnetic
performances [28]. The suggested machine, the HEFSM, has an efficiency ranging from
92 to 97 percent, and it is also used to dampen vibration and acoustic noise. Similar to
low-power, low-cost applications, single-phase hybrid SRMs are utilized for low-power,
low-cost applications. The new low-cost hybrid SRM for variable-speed pumps contains
permanent magnets in the shape of rectangular shards. As a result, both the average torque
and the torque ripple are increased; see [29].
However, the hybrid excitation method improves the efficiency of SRM due to the
strategic placement of PMs. The PM’s actions have the potential to deceive the SRM
because they alter the SRM’s typical behavior. Additionally, rare-earth metals and PMs
are becoming increasingly scarce. In the era of electric automobiles, both cost-effective
and efficient motors will be crucial. Nonetheless, the high demand and limited supply of
PMs will increase the cost of manufacturing PM-equipped machines. To maintain a high
torque/power density under these conditions, the PM’s electric motor integration strategy
is not a viable option. Therefore, this work primarily contributes to eliminating the need
for permanent magnets in SRMs, while simultaneously enhancing torque performance via
hybrid excitation techniques.

168
Energies 2022, 15, 7931

Due to the combined action of the two controllers, the proposed HESRM motor
has better control over the motor. Because separate controllers are used for the main
and auxiliary windings, it has more flexibility. The torque per ampere is improved by
the proposed method because it produces more torque at lower current levels than the
traditional SRM. The auxiliary windings can work at current control, while the main
windings can work in speed control. To increase torque, auxiliary windings can be used to
inject DC, and thus maximize the machine’s potential. There is less redundancy because
the motor is switched by two separate controllers, one of which can continue to operate the
machine even if the other fails. This machine exhibits a significantly higher flux in its stator
yoke than conventional machines, which is primarily caused by flux saturation in the core.
The remainder of this paper is organized as follows: Section 2 provides a detailed
description of the proposed HESRM’s machine topology and fundamental operations.
In Section 3, the simulation results of the proposed machine’s static and dynamic behavior,
including the results of the machine’s flux distribution, phase flux linkage, electromagnetic
torque, etc., are described in detail. Furthermore, Section 4 discusses the construction of
machines with identical dimensions and experimental validation. In Section 5, the average
torques of HESRM are compared with those of a conventional SRM. Finally, Sections 6 and 7
cover the discussion and conclusion in detail.

2. Machine Topology for the Proposed Machine


The machine topologies of conventional SRM and the proposed HESRM without PMs
are shown in Figure 1. Both motors have twelve stator and eight rotor poles without any
PMs incorporated within the cores. Still, the windings of poles per phase are assigned for
different excitation—main pole windings and auxiliary winding excitations, respectively,
for the assigned poles. The new hybrid excitation technique’s main goal is to boost the
conventional SRM’s torque performance without sacrificing any of its distinctive char-
acteristics. Permanent magnets and structural changes are entirely avoided in HESRM.
The excitation of the conventional SRM differs in that the windings are excited and sepa-
rated separately to inject more external current into the HESRM. These auxiliary windings
are selected so that two windings from each phase are separated, maintaining the motor’s
mechanical and electrical symmetry. Figure 1 depicts these alternately positioned auxiliary
windings, which inject external DC current in accordance with the torque required by the
controller. The motors’ measured speed is sensed and fed back to create PWMs, which are
then delivered to the throttle.

Figure 1. Machine topologies of two SRMs with the same rating as shown as (a) 12/8-pole conven-
tional SRM and (b) 12/8-pole HESRM without PMs.

First, the poles are assigned for auxiliary excitation. The stator of the HESRM differs
from that of the conventional SRM; in the HESRM, the stator poles are assigned to the main
and auxiliary poles based on the excitation. Since there are twelve stator poles in the motor,
six poles can be used for the primary excitation, while the remaining poles can be used for

169
Energies 2022, 15, 7931

the auxiliary excitation to provide the same three phases as conventional SRMs. Second,
the number of poles for the primary and auxiliary windings is maintained constant in
order to inject external flux into the poles simultaneously with the excitation of the primary
windings. Separate DC excitation can increase the resultant flux in the air gap due to the
symmetrical nature of the poles. Using electrical and mechanical symmetry, the primary
and secondary poles are differentiated. This geometry of twelve stators can be utilized
effectively to provide two distinct windings for each phase [30]. There are two distinct
windings in HESRM: the main windings and the auxiliary windings. With the aid of a
separately excited driver circuit, the auxiliary windings’ poles inject DC current. The main
windings provide the current for excitation of the motor. Because the variations in both
windings’ currents affect the air-gap flux density, the flux in HESRM is more than the
traditional SRM. The new topology has a greater flux density, power, and torque without
using the PM behavior.
Figure 2 is a structural diagram of magnetic field distribution that illustrates the
fundamental operating principle of the proposed HESRM without PMs. It shows the
magnetic field distribution of two SRMs with aligned and misaligned poles. Regardless
of whether the SRMs are operating in an aligned or unaligned position, identical field
lines can be observed. The HESRM flux is injected along the same path as the carryover
path for the primary winding flux. Therefore, the flux due to auxiliary excitation can
carry a greater flux to generate a greater electromagnetic torque. A detailed block diagram
of the HESRM is shown in Figure 3. The new hybrid excitation technique’s main goal
is to boost the conventional SRM’s torque performance without sacrificing any of its
distinctive characteristics. Permanent magnets and structural changes are entirely avoided
in HESRM. The excitation of the conventional SRM differs in that the windings are excited
and separated separately to inject more external current into the HESRM. These auxiliary
windings are selected so that two windings from each phase are separated, maintaining the
motor’s mechanical and electrical symmetry.
The net flux generated by HESRM in the absence of PMs travels through the core and
then the air gap, creating a closed path similar to the conventional flux path. When both
coils are excited, the flux intensity generated by the primary and secondary differs because
the secondary pole can carry a greater current than the primary. The total flux produced
by the primary coil and the auxiliary coils can be observed to be added. Consequently,
the total magnetic flux produced by the HESRM is greater than the conventional flux
without PMs. Consequently, the electromagnetic torque results from the algebraic sum of
the total effective flux of both windings.

Figure 2. Magnetic field distributions of two SRMs with 12/8 poles s shown as: (a) aligned pole and
(b) unaligned pole.

170
Energies 2022, 15, 7931

Figure 3. Block diagram of the proposed HESRM without permanent magnet excitation.

The dimensions and specifications of the conventional SRM and the proposed HESRM
are listed in Table 1. Similarly, the number of phases, number of poles in the stator and rotor,
rated power, speed, length of the core, and type of material are all identical. In addition,
the assignment of the HESRM’s auxiliary and primary poles differs based on the flux path.

Table 1. Main dimensions and parameters of two SRMs.

Dimensions Conventional SRM HESRM


Phase number 3 3
Stator and rotor poles 12/8 12/8
Hybrid excitation poles 12 6-6
Rated Power(kW) 1.2 1.2
Rated speed (rpm) 2500 2500
Outer dia of stator (mm) 138 138
Inner dia of rotor (mm) 81.50 81.50
Length of air gap (mm) 0.5 0.5
Type of material M15_26G M15_26G

3. Characteristics of the Proposed HESRM Using 2D FEA Analysis


This section presents the magnetic characteristics of the conventional SRM at aligned
and unaligned positions. To obtain better clarity and easiness of comparison, both the
SRMs are given the same size, number of coils per pole, and similar types of windings, as
shown in Table 1. The field lines where the stator and rotor iron have the shortest flux path
are concentrated in the aligned position rather than the unaligned one.

3.1. Magnetic Characteristics of Two SRM’s


The cross-sectional areas of the two SRMs are the same due to the same rotor con-
struction and geometric dimensions. Thus, both SRMs, the stators, and rotor iron weights
identical to the constructions have a 7.933 (kg) total net weight, and the stator and rotor
core steel consumptions (kg) are 7.05683 and 2.6735, respectively. Owing to the difference,
only the excitation scheme provided the winding with different currents.

Field Line and Flux Distributions


The magnetic characteristics of the proposed SRM are similar to that of the conven-
tional SRM. Here, the quantity of the flux lines will increase as more current is injected
into the auxiliary poles. As a result, the effective flux lines in the air gap increase, and
the resultant electromagnetic torque increases. Figures 4 and 5 demonstrate the magnetic
flux distribution and the flux density of the conventional SRM and the proposed HESRM,
respectively. The excitation for the three-phase scheme for the conventional SRM with

171
Energies 2022, 15, 7931

aligned and unaligned poles shows that the minimum flux density is 1.2–1.3 Tesla (T) for
the unaligned pole and 1.7–1.8 T for aligned poles. The conventional SRM’s range for maxi-
mum flux density is 0–1.9 T, and steel is the material used for the simulation. Figure 6 shows
the conventional SRM for the given excitation (5A) at aligned and unaligned positions.
At aligned position, it is 1.2 T, and at unaligned position, it is 1.6 T. Because the auxiliary
windings’ unexcited pole is aligned with the excitation pole in the HESRM, it can reach
1.06 T in alignment with the same excitation current (5A). Auxiliary injection current helps
to increase the magnetic flux density in an unaligned state in 1.3 T, comparing the magnetic
flux distribution and flux intensity of the proposed HESRM, where both characteristics
are identical. Accordingly, there is only one difference in the intensity of flux carrying in
the air gap due to different DC injections in the auxiliary poles. First, the excited state
produces the main flux with the effect of the main winding, and the extra DC provided by
the auxiliary excitation creates more electromagnetic torque.

Figure 4. Magnetic flux distribution of conventional SRM: (a) aligned poles and (b) unaligned poles.

Figure 5. Magnetic flux distribution of HESRM: (a) aligned poles and (b) unaligned poles.

The reluctance principle governs SRM. Therefore, the SRM generates the torque based
on the magnetic reluctance of the motor. The magnetic circuit between the rotor and stator
is very reluctant when they are not in alignment. As the rotor tries to align with the
powered stator poles at this point, the stator pole pairs are activated, reducing the magnetic
reluctance. Reluctance torque is formed when the rotor is able to reach the minimal point
of reluctance. It is necessary to precisely time the stator poles’ excitation so that it only
happens when the rotor is attempting to align with the exciting pole. SRM may require
positive feedback from encoders or Hall effect sensors in order to control the excitation of
the stator based on a precise rotor position.

172
Energies 2022, 15, 7931

Figure 6. Simulated flux density of conventional SRM and the hybrid DC-excited SRM.

3.2. Steady-State Characteristics


The electromagnetic behaviour of the proposed HESRM was analyzed through the
FEM (finite element method) using commercial software ANSYS/Maxwell. The motor drive
simulation was carried out by MATLAB/Simulink, and the interpretation was held by the
co-simulation method. For this purpose, the feasibility of the proof of concept was verified
with the help of software simulations by analyzing steady-state and dynamic behaviors.

Static Electromagnetic Torque


The static magnetic torque characteristics of two SRMs with phase excitations are
shown in Figure 7. The conventional SRM static torque curves with the current of 15 A vary
from 0 to 22.5 degrees, as shown in Figure 7a. Accordingly, the static torque for one-half of
the excitation cycle of conventional SRM attains unalignment to the aligned pole. It can be
seen that the torque varies from minimum to maximum and finally reaches a minimum
with the unaligned position. The maximum static torque that produces the phase excitation
current of 15 A is 2 N m. From the comparison, it is clear that the auxiliary excitation in
which a different current is provided with the help of DC injection may attain different
patterns of static currents. The auxiliary winding has a provision for carrying the maximum
current to the rated value. Figure 7b shows that the static flux curve pattern increases the
effect of varying according to the rises of DC in the auxiliary coils. Since additional current
is passed through DC injection in auxiliary windings using special circuitry, the flux inside
the core increases, which results in higher torque and power density.

Figure 7. Static torque characteristics: (a) conventional SRM at rated current 15 A. (b) HESRM under
various currents with respect to the DC field injection.

173
Energies 2022, 15, 7931

3.3. Dynamic Performance of Two SRM Drives


MATLAB/Simulink was used to study the performance of the motor drive system for
two SRMs. The proposed HESRM drive system’s Simulink model has an asymmetric half-
bridge converter, logic PWM, angle controller mechanical systems, DC sources, speed and
current controllers, a separate auxiliary excitation with an extra switch, etc. The machine
model is developed on the Ansys/Maxwell workbench. The dynamic performance of both
machines was performed and compared under the same conditions.

3.3.1. Gate Pulse Characteristics of Two SRMs


Figure 8 shows the moving torque and current characteristics of two SRMs in which the
gate pulses are given to the phases at the same turn ON and turn OFF times. In conventional
SRM, the single PWM pulses sequentially provide the switching scheme for all phases.
The hybrid excitation method provides identical gate pulses to the main and auxiliary
windings separately. The torque characteristics of conventional SRM can be achieved
with similar excitations in HESRM. Since the moving torque remains similar, the auxiliary
current can be further increased to the rated value to achieve maximum performance.

Figure 8. Comparison of characteristics of moving torque and current of two SRMs.

3.3.2. Closed-Loop Current Control of Two SRMs


The closed-loop current control of the conventional SRM is established by obtaining
current feedback from the windings and providing input to the current controller. The con-
ventional and proposed HESRM showed similar results at various input currents with
matching moving torque and slightly less ripple, as shown in Figure 9. The current from
the windings can be sensed and used to provide additional flux through the auxiliary in
this hybrid excitation method.

174
Energies 2022, 15, 7931

Figure 9. Comparison of closed-loop current control of conventional SRMs and HESRM.

3.3.3. Variable-Speed Control Strategy for Two SRMs


In Figure 10, the speed control is established using a proportional–integral (PI) con-
troller by measuring the actual speed and comparing it with the reference speed, which
varies continuously from minimum to rated value. The proportional (k p ) and integral
(k i ) values of the controller are tuned in such a way as to reach the set points as early as
possible. The analysis of the two results shows similar characteristics at the same current
rating and load. The additional torque needed for the motor can be identified and pro-
vided through these auxiliary windings to obtain better results. However, the HESRM
method is derived to enhance the performance by separately injecting the DC with the
help of auxiliary winding. Hence, these results elaborate on how to achieve the same
performance through the different winding excitations separately. However, the HESRM
makes it possible only through different currents passing through the assigned winding.
This method benefits by changing different currents through the controller to improve the
average torque characteristics and reduce the torque ripple to the minimum level.

Figure 10. Closed-loop variable-speed current control strategy: (a) conventional SRM and (b) HESRM;
the curves are current (A), torque (N-m), and speed (rpm).

175
Energies 2022, 15, 7931

In a conventional SRM, the variable speed is related to the main current excitation
and torque, whereas in a HESRM, where there are two currents—the main current and an
auxiliary current—the main current is responsible for both the excitations, and the auxiliary
speed current is responsible for the variable speed. Figure 10 displays the torque and
current characteristics of the traditional and hybrid SRM at variable speeds. To simulate
how EVs would operate in real time, the speed input is changed using ramps and steps.
Steps of 1000 rpm are used to ramp up the speed from zero to 2500 rpm. The graphs
show all three phases of the currents at various speeds. Since this motor is a three-phase
switching machine, the irregularities in the phase currents are caused by the speed changes.
The motor’s transient behavior is indicated by the longer switching patterns at the start of
the operation.

4. Experimental Results
4.1. Motor Prototype
The torque characteristic of the SRM is the most important performance parameter
that must be measured. As shown in Figure 11, in order to analyze the characteristics,
a back-to-back motor setup with a torque sensor coupled in between was constructed.
The motor used on the load side is a Permanent Magnet Synchronous Machine (PMSM)
rated at 5000 revolutions per minute (rpm). This motor is tightly coupled with the motor
of the plant. The non-linearity of the Futek torque sensor (model: TRD605) is +/−0.2%.
This high-precision strain gauge sensor measures up to 250 N-m. The load-side PMSM
contains an incremental speed encoder with a pulse per revolution (PPR) of 1024, which
can be increased to 5000 by counting the rising and falling edges of the encoder’s pulses.
The PMSM’s back EMF is used to drive a three-phase rectifier to a resistive load. Changes
to the resistance can be made to load the motor.

Figure 11. (a) Experimental setup of static measurements and (b) controller test setup of the pro-
posed HESRM.

4.2. Dynamic Performance of HESRM without PMs


The dynamic performance of the conventional SRM and the proposed HESRM can be
analyzed by enabling a closed-loop model. The experimental setup for both SRM drives
is identical in all aspects and conditions. For the analysis, conventional drive systems are
used, while in the auxiliary, additional circuits provide excitation. The motor has three hall
sensors that are electrically separated by one degree. Using an asymmetric inverter stack,
the three phases can be changed to run the motor. The control mechanism was enabled
using Wavect, a universal rapid control prototyping platform for Motor Control Drives
developed by Entuple Technologies, India. This FPGA-based controller implementation
was developed to analyze real-time strategies that can be modeled, tested, and managed.
The current sensors in the “wavect controller” can obtain the instantaneous phase currents,
and torque can be compared via an ADC channel in the controller. The primary advantage
of this mechanism is the ability to compare current and torque simultaneously. Figure 12
depicts the rewinding structure and the waveform corresponding to the obtained moving
torque and peak current when the motor is tested at various speeds and loaded to its
peak current.

176
Energies 2022, 15, 7931

Figure 12. Photograph of rewinding the conventional SRM to HESRM (a) Main winding (b) Auxi-
lary winding.

Figure 13 depicts the conventional SRM phase current and torque performance at
1000 and 2000 rpm. Figure 14 depicts the analysis of the auxiliary current and torque for
the hybrid excitation method. The three-phase current is denoted by different colours,
and green indicates the auxiliary exciting current. In this method, both the primary and
secondary poles are excited simultaneously without injecting any currents; the performance
matches the Ansys/Matlab software simulation results.

Figure 13. Conventional SRM dynamic performance of variable speed: (a) 1000 rpm and (b) 2500 rpm.

Figure 14. Proposed HESRM dynamic performance of variable speed: (a) 1000 rpm and (b) 2500 rpm.

The transient start-up response of the torque, phase current, and speed for two SRM
drives with closed-loop control are tested and shown in Figure 15. The results under
the same condition were evaluated with the DC-link voltage at 60 V and the load torque
at 2 N m, respectively. It was found that both SRMs perform start-up in a similar way,
according to the switching scheme provided, and accelerate rapidly as per the commanded
speed of 1000 rpm to rated 2500 rpm. They reach the time of 50 ms and 500 ms, respectively.
In other words, the speed changes from 500 rpm to 2500 rpm under closed-loop control,
and the gain parameter changes in a real-time manner to reach the commanded speed.
The result shows the rotor speed tracks the command signal in all manner and achieves the
substantial results obtained from the simulated one.

177
Energies 2022, 15, 7931

Figure 15. Closed-loop transient response at reference speed: (a) conventional SRM and (b) HESRM.

Figure 16 depicts the comparison between the measured and simulated average torque
characteristics. The graphical analysis shows that the measured and simulated torque
values are nearly identical. The average torque increased between 1000 and 2500 revolutions
per minute. In Figure 16a, at an initial speed of 1000 rpm, the average torque begins at
1.8 N m. It increases exponentially as the effect of the auxiliary current varies. Variable-
speed motors are permitted to adjust the auxiliary current from minimum to maximum,
up to the rated value. Therefore, there is a tendency to provide more flux, thereby increasing
the probability that the electromagnetic torque in the active region will increase.

Figure 16. Average torque performance for variable speed of proposed HESRM: (a) 1000, (b) 1500,
(c) 2000, and (d) 2500.

178
Energies 2022, 15, 7931

5. Comparison of Hardware Solution with FEM Simulation Results


In order to validate the proof of concept with results from the software analysis and
theoretical perdition, a 12/8 conventional SRM and a rewound SRM of the same size for the
proposed method were prototyped and experimentally validated. The main dimensions
and parameter specifications of the two SRMs are shown in Table 1.
Figure 17 depicts the comparison of the average torque between the conventional
SRM and the proposed HESRM at various speeds. To demonstrate the benefit of this novel
machine structure, the hybrid excitation without a permanent magnet is compared with
two SRMs from the same platform. The average performances for the speed range vary
from 1000 rpm to 2500 rpm, shown in different figures. At 1000 rpm, the torque is 2 N m
with the conventional current and excitation, while it reaches 2.5 N m with the auxiliary.
For this reason, there is a significant increase in torque performance, and the percentage of
torque increase at all speeds is greater than 20%. The pattern of torque response increase
varies depending on speed and load.

Figure 17. Comparison of average torque with variable speed for conventional SRM and proposed
HESRM: (a) 1000, (b) 1500, (c) 2000, and (d) 2500 rpm.

6. Discussion
Figure 18a shows the plots of conventional and HESRM’s average torque characteris-
tics for various speeds. Both have a difference of 1 to 2 N m at 500 rpm, and it gradually
increases. The HESRM reaches almost 2.4 N m at 1500 rpm, whereas the conventional one
is only 2.1 N m. The HESRM reaches a maximum torque of 2.6 N m at 2500 rpm with a
lead of almost 0.3 N m.

179
Energies 2022, 15, 7931

Figure 18. Comparison of conventional SRM with proposed HESRM: (a) average torque, (b) efficiency,
and (c) input power.

The efficiencies for both SRMs were almost identical at the beginning of the test at
500 rpm. As speed increases, the efficiency of the HESRM makes clear leads compared with
the conventional one. Up to 1500 rpm, the efficiencies of both motors follow a linear path
with a difference of approximately 10%. At the rated speed of 2500 rpm, the efficiency of
the motor differs to a maximum value of 80% and a minimum of 65%. The efficiency of the
HESRM increased by 20%, as shown in Figure 18b.
The input power consumed by both motors is almost constant at 500 rpm, as clearly
seen in Figure 18c. As the speed increases, the input power consumed by the motors starts
to increase. The conventional SRM intakes more power at all the respective rpm compared
with the proposed one. At 1000 rpm, the conventional SRM draws almost 850 W, whereas
the HESRM draws a minimum of 700 W. At maximum speed, the input power consumed
has a difference of almost 50 W.
Torque ripple is one of the major drawbacks of SRM, and thereby the hybrid excitation
methods tend to provide higher possibilities. The proposed new HESRM without PM tends
to exhibit less torque ripple than the conventional HESRM. In contrast, the conventional
SRM exhibits a ripple of 3.72% at 2500 rpm, which decreases to 3.69% at 2000 rpm. Simulta-
neously, the proposed HESRM reduces torque ripple as a result of reduced pole switching,
achieving 1.44% and 1.69% at 2500 and 2000 rpm, respectively.

7. Conclusions
When the average, static, and speed–torque characteristics of conventional and DC-
injected hybrid SRMs are compared, it was found that significant improvements were made
by HESRM over conventional SRM. In addition, the properties and performance of the
machines are improved without sacrificing their fundamental qualities. In addition to a
high torque-per-ampere rating, it offers high efficiency, high dependability, and redundancy.

180
Energies 2022, 15, 7931

The standard SRM is outperformed by the proposed method, and it generates greater torque
at significantly lower current levels. Furthermore, the proposed HESRM employs separate
excitations for the windings, thereby enhancing the versatility of the controller.
The proposed topology may be a replacement for HESRM machines with PM insertion.
Since this study clearly demonstrates enhanced performance, the proposed system can be
implemented in a variety of applications, such as electric vehicles. Moreover, improved
average and smooth speed–torque characteristics have been achieved, which can be an
excellent solution for EVs. Finally, this new topology can achieve greater efficiency with
reduced torque ripple. Future objectives of this research include the design and implemen-
tation of advanced, optimized controllers to further reduce torque ripples. We are planning
to study variable-speed control strategies for SRMs along with new designs.

Author Contributions: Conceptualization, V.A.; formal analysis, V.A. and M.J.H.; investigation, V.A.
and P.A.S.; writing—original draft preparation, V.A.; writing—review and editing, M.J.H. and G.L.;
supervision, M.J.H. and G.L.; industrial support, S.V.R. and T.P.M., concept analysis M.J.H. and S.V.R.;
experimental analysis, V.A. and P.A.S. All authors have read and agreed to the published version of
the manuscript.
Funding: The research is funded by the “University of Technology, Sydney”.
Data Availability Statement: Not applicable.
Acknowledgments: I would like to acknowledge “Entuple Technologies” for providing the universal
controller ‘Wavect 300’ and “Entuple E Mobility”, Bangalore, India, for supporting the entire hardware
implementation and testing setup during this work.
Conflicts of Interest: The authors declare no conflict of interest.

Abbreviations
The following abbreviations are used in this manuscript:

SRM Switched Reluctance Motor


HESRM Hybrid Excitation of Switched Reluctance Motors
PM Permanent Magnet
EVs Electric Vehicles
BLDCMs Brushless DC Machines
ICE Internal Combustion Engine
FEM Finite Element Analysis

References
1. Lan, Y.; Benomar, Y.; Deepak, K.; Aksoz, A.; Baghdadi, M.E. Switched Reluctance Motors and Drive Systems for Electric Vehicle
Powertrains: State of the Art Analysis andFuture Trends. Energies 2021, 14, 2079. [CrossRef]
2. Xu, J.; Zhang, L.; Meng, D.; Su, H. Simulation, Verification and Optimization Design of Electromagnetic Vibration and Noise of
Permanent Magnet Synchronous Motor for Vehicle. Energies 2022, 15, 5808. [CrossRef]
3. Men, X.; Guo, Y.; Wu, G.; Chen, S.; Shi, C. Implementation of an Improved Motor Control for Electric Vehicles. Energies 2022,
15, 4833. [CrossRef]
4. Martellucci, L.; Capata, R. High Performance Hybrid Vehicle Concept—Preliminary Study and Vehicle Packaging. Energies 2022,
15, 4025. [CrossRef]
5. Tomczewski, K.; Wrobel, K.; Rataj, D.; Trzmiel, G. A Switched Reluctance Motor Drive Controller Based on an FPGA Device with
a Complex PID Regulator. Energies 2021, 14, 1423. [CrossRef]
6. Rahman, M.S.; Lukman, G.F.; Hieu, P.T.; Jeong, K.; Ahn, J. Optimization and Characteristics Analysis of High Torque Density
12/8 Switched Reluctance Motor Using Metaheuristic Gray Wolf Optimization Algorithm. Energies 2021, 14, 2013. [CrossRef]
7. Pupadubsin, R.; Kachapornkul, S.; Jitkreeyarn, P.; Somsiri, P.; Tungpimolrut, K. Investigation of Torque Performance and Flux
Reversal Reduction of a Three-Phase 12/8 Switched Reluctance Motor Based on Winding Arrangement. Energies 2022, 15, 284.
[CrossRef]
8. Wang, J.; Jiang, W.; Wang, S.; Lu, J.; Williams, B.W.; Wang, Q. A Novel Step Current Excitation Control Method to Reduce the
Torque Ripple of Outer-Rotor Switched Reluctance Motors. Energies 2022, 15, 2852. [CrossRef]
9. Kocan, S.; Rafajdus, P.; Bastovansky, R.; Lenhard, R.; Stano, M. Design and Optimization of a High-Speed Switched Reluctance
Motor. Energies 2021, 14, 6733. [CrossRef]

181
Energies 2022, 15, 7931

10. Ding, W.; Member; Liu, G.; Li, P. A hybrid control strategy of hybrid-excitation switched reluctance motor for torque ripple
reduction and constant power extension. IEEE Trans. Ind. Electron. 2020, 67, 38–48. [CrossRef]
11. Kabir, M.A.; Husain, I. Design of Mutually Coupled Switched Reluctance Motors (MCSRMs) for Extended Speed Applications
Using 3-Phase Standard Inverters. IEEE Trans. Energy Convers. 2016, 31, 436–445. [CrossRef]
12. Elmutalab, M.A.; Elrayyah, A.; Husain, T.; Sozer, Y. Extending the Speed Range of a Switched Reluctance Motor Using a Fast
Demagnetizing Technique. IEEE Trans. Ind. Appl. 2018, 54, 3294–3304. [CrossRef]
13. Lee, C.; Krishnan, R.; Lobo, N.S. Novel Two-Phase Switched Reluctance Machine Using Common-Pole E-Core Structure: Concept,
Analysis, and Experimental Verification. IEEE Trans. Ind. Appl. 2009, 45, 703–711. [CrossRef]
14. Mecrow, B.C.; El-Kharashi, E.A.; Finch, J.W.; Jack, A.G. Segmental rotor switched reluctance motors with single-tooth windings.
IEE-Electr. Power Appl. 2003, 150, 591–599. [CrossRef]
15. Mecrow, B.C.; El-Kharashi, E.A.; Finch, J.W.; Jack, A.G. Preliminary Performance Evaluation of Switched Reluctance Motors With
Segmental Rotors. IEEE Trans. Energy Convers. 2004, 19, 679–686. [CrossRef]
16. R, V.; Fernandes, B.G. Design Methodology for High-Performance Segmented Rotor Switched Reluctance Motors. IEEE Trans.
Energy Convers. 2015, 30, 11–21. [CrossRef]
17. Ding, W.; Yang, S.; Hu, Y.; Li, S.; Wang, T.; Yin, Z. Design Consideration and Evaluation of a 12/8 High-Torque Modular-Stator
Hybrid Excitation Switched Reluctance Machine for EV Applications. IEEE Trans. Ind. Electron. 2017, 64, 9221–9232. [CrossRef]
18. Ding, W.; Fu, H.; Hu, Y. Characteristics Assessment and Comparative Study of a Segmented-Stator Permanent-Magnet Hybrid-
Excitation SRM Drive With High-Torque Capability. IEEE Trans. Power Electron. 2018, 33, 482–500. [CrossRef]
19. Hasegawa, Y.; Nakamura, K.; Ichinokura, O. A Novel Switched Reluctance Motor With the Auxiliary Windings and Permanent
Magnets. IEEE Trans. Magn. 2012, 48, 3855–3858. [CrossRef]
20. Zhihui, C.; Yaping, S.; Yangguang, Y. Static Characteristics of a Novel Hybrid Excitation Doubly Salient Machine. In Proceedings
of the International Conference on Electrical Machines and Systems (ICEMS), Nanjing, China, 27–29 September 2005. [CrossRef]
21. Sulaiman, E.; Kosaka, T.; Matsui, N. High Power Density Design of 6-Slot–8-Pole Hybrid Excitation Flux Switching Machine for
Hybrid Electric Vehicles. IEEE Trans. Magn. 2011, 47, 4453–4456. [CrossRef]
22. Jia, S.; Qu, R.; Li, J.; Li, D.; Kong, W. A Stator-PM Consequent-Pole Vernier Machine With Hybrid Excitation and DC-Biased
Sinusoidal Current. IEEE Trans. Magn. 2017, 53, 1–4. [CrossRef]
23. Ahn, J.-W.; Park, S.-J.; Lee, D.-H. Hybrid Excitation of SRM for Reduction of Vibration and Acoustic Noise. IEEE Trans. Ind.
Electron. 2004, 51, 374–380. [CrossRef]
24. Lu, K.; Jakobsen, U.; Rasmussen, P.O. Single-Phase Hybrid Switched Reluctance Motor for Low-Power Low-Cost Applications.
IEEE Trans. Magn. 2011, 47, 3288–3291. [CrossRef]
25. Lu, K.; Rasmussen, P.O.; Watkins, S.J.; Blaabjerg, F. A New Low-Cost Hybrid Switched Reluctance Motor for Adjustable-Speed
Pump Applications. IEEE Trans. Ind. Appl. 2011, 47, 314–321. [CrossRef]
26. Lu, K.Y.; Rasmussen, P.O.; Watkins, S.J.; Blaabjerg, F. A New Low-Cost Hybrid Switched Reluctance Motor for Adjustable-Speed
Pump Applications. In Proceedings of the 2006 IEEE Industry Applications Conference Forty-First IAS Annual Meeting, Tampa,
FL, USA, 8–12 October 2006. [CrossRef]
27. Nakamura, K.; Ichinokura, O. Super-MultipolarPermanent Magnet Reluctance Generator Designed for Small-Scale Wind-Turbine
Generation. IEEE Trans. Magn. 2012, 48, 3311–3314. [CrossRef]
28. Nakamura, K.; Ichinokura, O. Super-Multipolar Permanent Magnet Reluctance Generator Designed for Small-Scale Renew-
able Energy Generation. In Proceedings of the International Conference on Electrical Machines, ICEM, Marseille, France,
2–5 September 2012. [CrossRef]
29. Lee, C.; Krishnan, R. New Designs of a Two-Phase E-Core Switched Reluctance Machine by Optimizing the Magnetic Structure
for a Specific Application: Concept, Design, and Analysis. IEEE Trans. Ind. Appl. 2009, 45, 1804–1814. [CrossRef]
30. Abhijith, V.; Hossain, M.J.; Lei, G.; A S, P. Performance Improvement of Switched Reluctance Motor Using Hybrid Excitation
Method Without Permanent Magnets. In Proceedings of the 2021 24th International Conference on Electrical Machines and
Systems (ICEMS), Gyeongju, Korea, 31 October–3 November 2021.

182
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
Tel. +41 61 683 77 34
Fax +41 61 302 89 18
www.mdpi.com

Energies Editorial Office


E-mail: [email protected]
www.mdpi.com/journal/energies
Academic Open
Access Publishing

www.mdpi.com ISBN 978-3-0365-7657-2

You might also like