Fulltext

Download as pdf or txt
Download as pdf or txt
You are on page 1of 219

THESIS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

Inverter Design for SiC-based Electric Drive


Systems with Optimal Redundant States Control
of Space Vector Modulation

GEORGIOS MADEMLIS

Department of Electrical Engineering


Division of Electric Power Engineering
CHALMERS UNIVERSITY OF TECHNOLOGY
Gothenburg, Sweden 2021
Inverter Design for SiC-based Electric Drive Systems with Optimal Redundant
States Control of Space Vector Modulation
GEORGIOS MADEMLIS
ISBN 978-91-7905-443-4

Revised March 8, 2021

© GEORGIOS MADEMLIS, 2021.

Doktorsavhandlingar vid Chalmers Tekniska Högskola


Ny serie nr. 4910
ISSN 0346-718X

Electrical Machines and Power Electronics


Division of Electric Power Engineering
Department of Electrical Engineering
Chalmers University of Technology
SE–412 96 Göteborg
Sweden
Telephone +46 (0)31-772 1000

Chalmers Bibliotek, Reproservice


Gothenburg, Sweden 2021
To my family and friends
Inverter Design for SiC-based Electric Drive Systems with Optimal Redundant
States Control of Space Vector Modulation
GEORGIOS MADEMLIS
Department of Electrical Engineering
Chalmers University of Technology

Abstract
The need for inverters with ever increasing power density and efficiency has
recently become the driving factor for research in various fields. Increasing the
operating voltage of the whole drive system and utilizing newly developed SiC
power switches can contribute towards this goal. Higher operating voltage allows
the design of drives with lower current, which leads to lower copper losses in cables
and machine, while SiC switches can drastically increase the inverter efficiency.
Offshore renewable power generation, such as tidal power, is a typical application
where the increase of operating voltage can be highly beneficial. The ongoing
electrification of transportation calls also for high power electric powertrains with
high power density, where SiC technology has key advantages.
In the first part of the thesis, suitable control schemes for inverters in synchronous
machine drive systems are derived. A properly designed Maximum Power Point
Tracking algorithm for kite-based tidal power systems is presented. The speed
and torque of this new tidal power generation system varies periodically and the
inverter control needs to be able to handle this variable power profile. Experimental
verification of the developed control is conducted on a 35 kVA laboratory emulator
of the tidal power generation unit.
Electric drives using multilevel inverters are studied afterwards. Multilevel inverters
use multiple low-voltage-rated switches and can operate at higher voltage than
standard two-level inverters. The Neutral Point Clamped (NPC) converter is a
commonly used multilevel inverter topology for medium voltage machine drives.
However, the voltage balancing of its dc-side capacitors and the complexity of
its control are still issues that have not been effectively solved. A new method
for the optimal utilization of the redundant states in Space Vector pulse-width-
Modulation (SVM) is proposed in this thesis in order to control its dc-link voltages.
Experimental verification on a 4-kV-rated prototype medium-voltage PMSM drive
with 5-level NPC converters is conducted in order to validate the effectiveness of
the proposed control technique.

v
Low switching and conduction losses are typical characteristics of SiC switches
that can be utilized to build inverters with high power density, due to the increased
efficiency and smaller form-factor. Due to the above, SiC power modules have been
particularly attractive for the automotive industry. The design approach of 2-level
automotive inverters has been studied in this project. Moreover, a new design
approach for the cooling system of automotive inverters has been developed in
this thesis, which fine-tunes the inverter heatsink utilizing standard legislated test
routines for electric vehicles. Multiple conjugate-heat-transfer (CHT) computation
results showcase the iterative optimization procedure on a test-case 250 kW (450 A)
automotive SiC inverter.
Finally, the experimental testing of high power machine drives in order to verify
the control and the hardware design is an important step of the development pro-
cess. Thus, the performance of the prototype 450 A SiC 2-level inverter has been
experimentally validated with a power hardware-in-the-loop (P-HIL) set-up that
emulates an automotive drive system. Several challenges have been addressed with
respect to the accurate modelling of the motor and the control of the circulating
power in the system. A new control technique utilizing the redundant states of the
SVM has been developed for this set-up to effectively suppress the zero-sequence
circulating current to 3.3% of the line current at rated power.

Index Terms: Common Mode Voltage, Electric Vehicle, Maximum Power Point
Tracking (MPPT), Medium Voltage Converters, Multilevel, Space Vector Modula-
tion, Neutral Point Clamped (NPC), Power Hardware-in-the-Loop, Silicon Carbide
(SiC), Tidal power, Voltage balancing.
Acknowledgements
Firstly, I would like to express my gratitude to my supervisor Prof. Yujing Liu for
his guidance and support throughout the course of this PhD project. Thank you
very much for believing in my ideas and encouraging me to investigate the areas
that I enjoyed mostly and at the same time for providing me with comments that
helped me to improve myself as a person and as a future researcher. I thank also
my examiner Prof. Ola Carlson and my co-supervisor Dr. Stefan Lundberg for
planning my PhD studies and for their feedback on my thesis.
I am grateful for the support I received from my colleagues in the Electric Power
Engineering division. I would like to thank Nimananda Sharma, Dr. Raik Orbay
and Assoc. Prof. Peiyuan Chen for our collaboration and the fun time we spent
working together in our common projects. I thank also Junfei Tang who has
always been willing to provide support and help dedicating so much of his time.
Many thanks to Ankur, Christos, Kyriaki and Artem for the amazing time we
spent together all these years.
This work has been carried out within the PowerKite project during 2016-2019
(European Union’s Horizon 2020 - grand agreement No 654438) and within the
iTEM project during 2020-2021 (supported by the Swedish Energy Agency). Their
financial contribution is highly appreciated.
Finally, I would like to thank my partner Ioanna Teknetzi and my family for
always being by my side.

Γώργιoς Mαδµλής / Georgios Mademlis


Gothenburg, Sweden
December, 2020

vii
List of Acronyms

2LC 2-level Converter


ADC Analog-to-Digital Converter
ANPC Active Neutral Point Clamped
B2B Back-to-Back
BEV Battery Electric Vehicle
CAD Computer-aided Design
CFD Computational Fluid Dynamics
CHT Conjugate Heat Transfer
DBC Direct Bond Copper
DSP Digital Signal Processor
DTC Direct Torque Control
EESM Electrically Excited Synchronous Machine
EMF Electromotive Force
EMI Electromagnetic Interference
ePWM enhanced Pulse Width Modulator
ESL Equivalent Series Inductance
ESR Equivalent Series Resistance
FFT Fast Fourier Transform
FC Flying Capacitor
FEM Finite Element Method
FOC Field Oriented Control
FPGA Field-Programmable Gate Array
GaN Gallium Nitride
HVDC High Voltage Direct Current
IC Integrated Circuit

ix
IGBT Insulated-Gate Bipolar Transistor
IM Induction Machine
IPMSM Interior Permanent Magnet Synchronous Machine
IUT Inverter-Under-Test
LPF Low-Pass-Filter
L2L Line-to-Line
MCU Microcontroller Unit
MMC Modular-Multilevel-Converter
MOSFET Metal Oxide Semiconductor Field Effect Transistor
MPPT Maximum Power Point Tracking
MTPA Maximum Torque Per Ampere
NPC Neutral Point Clamped
PCB Printed Circuit Board
PCC Point of Common Coupling
PI Proportional-plus-Integral
PLL Phase-Locked Loop
PMSG Permanent Magnet Synchronous Generator
PMSM Permanent Magnet Synchronous Machine
pu Per Unit
PWM Pulse-Width Modulation
RES Renewable Energy Sources
rms root mean square
rpm revolutions per minute
Si Silicon
SiC Silicon Carbide
SG Synchronous Generator
SMD Surface-Mount Device
STATCOM Static synchronous Compensator
SPWM Sinusoidal Pulse-Width Modulation
SVM Space Vector Modulation
THD Total Harmonic Distortion
TSR Tip Speed Ratio
TUSK Tethered UnderSea Kite
VSC Voltage Source Converter
WBG Wide-Bandgap
List of Symbols
Each variable symbol has italic font and is defined in the text where it is first
mentioned. Bold symbols in the thesis describe vectors.
The basic variables of the thesis defined below are sorted alphabetically:

reference signal
θ angle [°]
τ time constant [s]
ωr angular velocity of the electrical machine [rad/s]
ωturb rotational speed of the turbine [rad/s]
ac machine current controller bandwidth [rad/s]
aω generator speed controller bandwidth [rad/s]
acs grid current controller bandwidth [rad/s]
aP LL grid PLL bandwidth [rad/s]
aDClink grid dc-link controller bandwidth [rad/s]
Ba active damping term [Ω]
CP performance coefficient of the tidal turbine [%]
D altitude angle of the TUSK plant [deg] (only for Chapter 3)
dT duty cycle of the power switch T
fsw switching frequency [Hz]
i trajectory position of the kite (only for Chapter 3)
i0 zero-sequence current [A]
ibl ac current used for voltage balancing of the 5-level NPC drive [A]
Id1 ... Id5 dc currents at the clamping points of a 5-level NPC converter [A]
ie motor stator current [A]
igen generator stator current [A]
is grid current [A]
J moment of inertia [kg·m2 ]
KI integral gain of a PI controller
KP proportional gain of a PI controller

xi
Ld stator inductance in the d-axis [H]
Lf grid filter inductance [H]
Lq stator inductance in the q-axis [H]
L1 ... L3 inductances of the balancing converter for the NPC drive [H]
m̂ modulation index [%]
N number of levels of the phase voltage of a VSC
p electrical machine’s pole pairs
~ x, y, z parametric equation of the Viviani’s Curve in space coordinates

P
Plossi power losses at the trajectory point i owed to the drag of the kite [W]
Pm turbine mechanical power [W]
Pgrid active power at the PCC [W]
Qgrid reactive power at the PCC [W]
r cylinder radium of the Viviani’s Curve [m] (only for Chapter 3)
R sphere radium of the Viviani’s Curve [m] (only for Chapter 3)
RS machine stator resistance [Ω]
RT turbine radius [m]
t time [s]
Te electromagnetic torque of the machine [Nm]
Tgrid electrical torque at the PCC [Nm]
Tm mechanical torque of the machine [Nm]
tr rise time [s]
Tsw switching period of a power converter [s]
tD ... tG dwell times of the switching states in the SVM [s]
v∗ normalized reference voltage of a VSC
v0 common mode voltage [V]
Va effective water speed [m/s]
VC1 ... VC4 dc capacitor voltages of a 5-level NPC converter [V]
VCC supply voltage for electronic circuits [V]
Vdc voltage of each dc-link capacitor bank [V]
Vdc tot total dc-link voltage of a drive [V]
VD ... VG space vectors of the SVM in (α0 , β 0 ) coordinates
VD0 ... VG0 three-phase switching vectors of the SVM
ve machine terminal voltage [V]
VGS MOSFET gate-source voltage [V]
vs grid voltage at the PCC [V]
Vtide tidal free stream speed [m/s]
Vtr voltage of the triangular carrier in SPWM [V]

V̂ph reference of the amplitude of the converter phase voltage [V]
Contents

Abstract v

Acknowledgements vii

Contents xiii

1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Motor Drives in Battery Electric Vehicles . . . . . . . . . . 2
Electric Motor Selection . . . . . . . . . . . . . . . . . . . . 3
Inverter Design Considerations . . . . . . . . . . . . . . . . 4
Wide-Bandgap Technology in Electromobility . . . . . . . . 5
1.1.2 Kite-based Tidal Power Generation System . . . . . . . . . 6
Optimal Control of Tidal Power Generator . . . . . . . . . 8
Medium Voltage Multilevel Machine Drives . . . . . . . . . 9
Wide-Bandgap Technology in Offshore Power Generation . 11
1.2 Aims of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.4 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.5 List of Publications . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2 Synchronous Machine Drives: Topologies and Control 17


2.1 Structure of Electrical Machine Drives . . . . . . . . . . . . . . . . 17
2.2 Field Oriented Control . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2.1 Speed Controller Design . . . . . . . . . . . . . . . . . . . . 19
2.2.2 Current Reference Generation . . . . . . . . . . . . . . . . . 20
2.2.3 Machine Current Controller Design . . . . . . . . . . . . . . 21
2.3 Vector Control for Grid-Side Converter . . . . . . . . . . . . . . . . 23

xiii
2.3.1 Real-/Reactive-Power Control Design . . . . . . . . . . . . 23
2.3.2 Grid-side Current Controller Design . . . . . . . . . . . . . 25

3 Modelling and Control of Kite-based Tidal Power Generator Sys-


tems 27
3.1 Introduction to Control Techniques of the Kite . . . . . . . . . . . 27
3.2 Tidal Power Plant System Modeling . . . . . . . . . . . . . . . . . 28
3.2.1 Modeling of Kite’s Trajectory and Power Generation . . . . 30
3.2.2 MPPT with Active Control of Generator Rotor Speed . . . 35
3.2.3 MPPT with Torque Control of the Generator . . . . . . . . 36
3.3 Design of Tidal Power Laboratory Emulator . . . . . . . . . . . . . 37
3.4 Experimental Results of the Laboratory Emulator . . . . . . . . . 40
3.4.1 Effect of the Speed Controller’s Response Time on the kite’s
Power Generation . . . . . . . . . . . . . . . . . . . . . . . 41
3.4.2 MPPT Algorithm Comparison Study . . . . . . . . . . . . . 45
3.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

4 Multilevel Converter Topologies and Modulation Techniques 53


4.1 Motivation towards Medium Voltage Drives . . . . . . . . . . . . . 53
4.2 Multilevel Converter Topologies & Capacitor Voltage Balancing . . 54
4.3 Multilevel PWM Modulation Strategies . . . . . . . . . . . . . . . 58
4.3.1 Sinusoidal PWM . . . . . . . . . . . . . . . . . . . . . . . . 58
4.3.2 Space Vector Modulation . . . . . . . . . . . . . . . . . . . 60
4.4 Electric Drive Design using 3-level NPC Converters . . . . . . . . . 67
4.5 Electric Drive Design using 5-level NPC Converters . . . . . . . . . 70
4.5.1 SVM-based DC-Link Voltage Balancing . . . . . . . . . . . 72
4.5.2 Hardware-based DC-Link Voltage Balancing . . . . . . . . . 81
4.5.3 Combined Hardware and SVM-based Voltage Balancing . . 89
4.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

5 Design and Testing of Medium Voltage PMSM Drive 97


5.1 System Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
5.2 Prototype PMSG Drive Design . . . . . . . . . . . . . . . . . . . . 98
5.2.1 SiC Gate Driver Design . . . . . . . . . . . . . . . . . . . . 101
5.2.2 NPC Converter Board . . . . . . . . . . . . . . . . . . . . . 103
5.2.3 DC-Link Board . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.2.4 Measurement Devices . . . . . . . . . . . . . . . . . . . . . 107
5.2.5 Overcurrent-Overvoltage Protection Strategy . . . . . . . . 109
5.3 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . 111
5.3.1 Medium Voltage PMSG . . . . . . . . . . . . . . . . . . . . 111
5.3.2 Power Conversion System . . . . . . . . . . . . . . . . . . . 114
5.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

6 Power-HIL Testing of High-Power Electrical Machine Drives 121


6.1 Electrical Machine Emulator . . . . . . . . . . . . . . . . . . . . . 121
6.1.1 Machine Modelling Approaches . . . . . . . . . . . . . . . . 122
6.1.2 PMSM Emulation with the Inverted Machine Model . . . . 124
6.2 Common Mode Voltage Control . . . . . . . . . . . . . . . . . . . . 125
6.2.1 Closed-loop Common Mode Current Control . . . . . . . . 127
6.2.2 Zero-Sequence Voltage Compensation . . . . . . . . . . . . 128
6.2.3 Operation close to the Voltage Limit . . . . . . . . . . . . . 129
6.3 Comparative Results of PMSM drive and Machine Emulator . . . 131
6.3.1 Experimental Results of a 60 kW PMSM Drive . . . . . . . 131
6.3.2 Experimental Results of PMSM Emulator . . . . . . . . . . 135
6.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

7 Electrical and Thermal Design of Automotive SiC Inverter 145


7.1 Design Specifications . . . . . . . . . . . . . . . . . . . . . . . . . . 145
7.1.1 Design Specifications for Light-Duty Vehicles . . . . . . . . 146
7.1.2 Design Specifications for Heavy-Duty Vehicles . . . . . . . . 147
7.2 SiC 2-Level Inverter Design Requirements . . . . . . . . . . . . . . 148
7.3 DC-Link Capacitor Bank Design . . . . . . . . . . . . . . . . . . . 150
7.4 Design of Liquid Cooling System for Inverters . . . . . . . . . . . . 153
7.4.1 Overview of Cooling Plate Geometries . . . . . . . . . . . . 154
7.4.2 CHT Computation Approach with Steady-State Heat Load 156
7.4.3 CHT Computation Process with Transient Heat Dissipation 159
7.4.4 Iterative Cooling Design Process . . . . . . . . . . . . . . . 165
7.5 Inverter Efficiency Measurements . . . . . . . . . . . . . . . . . . . 167
7.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168

8 Conclusions and Future Work 171


8.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
8.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173

References 175

Appendices 194
Chapter 1

Introduction

1.1 Background
Electrical machine drives with high efficiency and power density are currently
required in many applications and are under research by both the industry and
the academia. Electrification of ground transportation is seen nowadays as one of
the most viable solutions for sustainable transports. The electric drive, including
a motor and an inverter, is one of the main components of an electric vehicle. An
increase of the drive system efficiency is highly beneficial, since this could increase
the energy autonomy of battery-powered vehicles. At the same time, automotive
drive systems with higher power capabilities than in the past are now requested by
customers, while the industry strives to keep the volume, weight and cost of the
powertrain as low as possible. Therefore, there is a high demand towards electric
drives with high power density.
Exploitation of new renewable energy sources for electric power generation is
another application where further improvement of the efficiency and power density
of the generator drive has high importance. For example, harvesting electric power
from sea tides has seen growing research interest and, thus, the design of the power
conversion system is a highly important and challenging task.
The state of the art and the main challenges in designing machine drives for electric
vehicles and renewable energy source applications are presented in the following
two subsections.

1
Chapter 1. Introduction

1.1.1 Motor Drives in Battery Electric Vehicles


The propulsion system of a battery electric vehicle (BEV - i.e. vehicle that uses
rechargeable batteries as its main energy source) has the following typical structure
as cited from [1] and shown in Fig. 1.1. The main motor drive consists of an
electric motor and a three-phase inverter. The dc-side of the inverter is supplied
by the high-voltage (HV) battery, while a low-voltage (LV) battery is utilized to
provide energy to the low-power electronics of the vehicle. The structure of the
vehicle charging system and charging ports are described by the IEC 61851-23:2014
and IEC 62196-3:2014 standards [2, 3] and by SAE J1772 [4]. The vehicle has an
on-board charger suitable for charging the HV battery through the ac grid, while
dc charging is also possible to enable fast charging capabilities.
Modifications of the standard electric powertrain structure have been suggested to
reduce the number of components and the cost. For example, the inverter of the
main powertrain can be utilized also as an on-board charger [1]. Wireless charging
of the HV battery becomes also more and more attractive as this technology
evolves [5]. A dc/dc converter between the HV battery and the dc-link of the drive
has been used for the cases when the variable voltage of the HV battery does not
match the rated voltage of the drive [6]. Hybrid electric vehicles have also similar
powertrain structure, with the addition of the internal combustion engine, which
can be placed in parallel or in series to the electric powertrain [7].
A closed-loop cooling system is installed in the vehicle that is responsible for
extracting the heat out of the motor, the inverter and the on-board charger.

HV
Motor drive battery LV
gearbox auxiliary battery

Heat
exch. Pump

charger

or
electrical connections DC + - Grid
liquid coolant flow charger (3ph or 1ph)
mechanical connections

Fig. 1.1: Schematic diagram of the powertrain in a battery electric vehicle and its basic
cooling system

2
1.1. Background

Variations of that system exist depending on the type of the coolant [8], which can
be oil, water with ethylene glucole or a combination of these two. The main cooling
loop studied in this thesis is shown in Fig. 1.1, where the power converters and
the electric motor are connected in series with the pump and the heat exchanger.
A secondary cooling loop can also cool down the battery pack, as suggested in [9].

Electric Motor Selection

The design of the electric motor is one of the main fields of study related with
electromobility. The final selection of the machine type is determined by the
specifications of the vehicle and the application where it is going to be used.
Electrical machines with permanent magnets are the dominant machine types in
electric vehicles and a list of common design solutions have been reported in [10].
The interior permanent magnet synchronous machine (IPMSM) is widely used
when high torque density and better efficiency at low and medium speeds are
required [11].
Alternative machines without permanent magnets have also been used in electric
vehicles [12]. The induction machine (IM) is a common machine type, which has
typically lower torque density and lower efficiency than the PMSM, especially at
high-torque low-speed operation [13]. The switched reluctance machine (SRM) is
also an alternative machine type, which is mostly suitable for high-speed operation
[11]. Permanent-magnet-free synchronous machines have recently seen research
interest for automotive applications, such as the electrically excited synchronous
machine (EESM) [14, 15], which has been studied in comparison with the IPMSM
during this PhD project. Many other design aspects of electric motors for traction
applications are currently under research, but are not further investigated here. For
example, there are studies related with the winding distribution type for IPMSM,
such as in [16, 17] where distributed windings are compared with fractional-slot
windings. The optimal stator winding types for automotive machines have also
been investigated in [18], where hairpin and random windings have been compared.
Therefore, it is not easy to draw a conclusion about the optimal machine design for
an electric vehicle, since many aspects need to be taken into consideration. The
speed range of the motor can also be selected in correlation with the transmission
system design [19, 20], which can be of single-gear or have multiple gears.
Powertrains with multiple motors have been suggested for heavy-duty vehicles,
such as dual-motor designs for buses [21] and large trucks [22] and can also be
found in passenger vehicles, such as in Tesla BEVs [23]. In-wheel direct-driven

3
Chapter 1. Introduction

motor drives have also been suggested for passenger BEVs with two-wheel drives
[24] or even four-wheel drives [25], where each drive has its individual motor and
inverter. Special design requirements exist for the motors of these vehicles.

Inverter Design Considerations

The main inverter of the vehicle that drives the motor is another key component
of the powertrain. It is responsible for converting the dc voltage of the battery
into ac voltage with the correct frequency and amplitude for the motor. The
three-phase voltage source converter (VSC) is the standard inverter type used in
the automotive industry [26] and, therefore, it is selected for this study. The main
design requirements of the inverter is to have high power density and efficiency,
while the cost of the system is kept as low as possible and the vehicle specifications
in terms of maximum power, current and voltage are satisfied. Several design
aspects can be investigated to optimize the design based on these requirements.
The type of semiconductors used for the power switches and their rated voltage,
as well as the inverter topology are some of these research aspects.
Based on current industry trends, the majority of the motor drive ratings range
between 30 kW and 200 kW [1], while the majority of traction inverters in the
future will be in the 100-500 kW range judging from the ongoing research and
industry practices [27]. Newly released passenger BEVs with similar price such as
Volvo XC40 Recharge [28], Tesla Model 3 [23], Polestar 2 [29] and the expected
for 2021 BMW i4 concept car [30] have total power of their powertrain between
300 and 390 kW.
The rated operating voltage of the drive depends on the high-voltage battery,
which is currently rated to 400 V in the majority of the cases. However, the need
for fast-charging capabilities of the newer BEV models may also be accompanied
with a higher-voltage battery at 800 V increasing at the same time the rated
voltage of the motor drive [27, 31, 32]. The use of higher dc voltage will also reduce
of the current of the drive and, also, the copper losses of the motor.
Because of this trend towards higher voltage in automotive drives, multilevel
inverters and specifically 3-level NPC converters have been considered as an
alternative to the 2-level converter (2LC). As will be discussed later in this chapter,
these inverters have much lower Total Harmonic Distortion (THD) than the 2-level
topology for the same switching frequency [33], which will also have an impact on
the stator losses of the motor, and are capable of handling even higher voltages
than 800 V. However, since the current trend in the automotive industry is to have

4
1.1. Background

maximum 800 V system, the high-voltage capabilities of the multilevel converters


cannot be utilized. These inverters suffer also from higher cost and complexity
due to their larger number of components [27]. In case higher-voltage voltage
automotive drives are considered in the future, multilevel inverters could be an
attractive candidate.

Wide-Bandgap Technology in Electromobility

Wide-bandgap (WBG) semiconductors have recently evolved into an attractive


solution for power converters of electric vehicles due to the following distinctive
advantages compared to silicon (Si) counterparts. Silicon Carbide (SiC) and
Gallium Nitride (GaN) are the two most industrialized WBG semiconductors. As
evident from the name of these materials, they have a larger bandgap compared to
Si. Higher bandgap means that semiconductor dies can withstand larger voltage
compared with Si dies of the same thickness. On the other hand, if the voltage
rating is kept the same, the semiconductor die can have smaller thickness, higher
doping is possible and the resultant on-resistance can be much lower. WBG
devices can also have faster switching transients and the switching losses can be
extremely low, even at high voltage, that allows the development of faster-switching
converters with lower THD. Higher operating temperatures are also possible for
WBG semiconductors, even higher than 250°C, and the main limitation is basically
the temperature constraints of the packaging.
As a matter of fact, SiC has more than 10 times higher breakdown field compared
to Si [34], which leads to 10 times thinner SiC devices of same voltage rating
with the Si device. The loss reduction is even larger and when combined with
the higher operating temperature, the power density of the final converter can be
substantially enhanced.
SiC MOSFETs are competing mainly with Si IGBTs in applications where the
commonly found 1200 V switches are used [34]. Therefore, they are a suitable
candidate for the main inverter of a BEV with rated dc voltage of 800 V. Automotive
inverters with Si IGBTs are still capable of operating with 800 V dc-link, but they
have significantly higher switching losses [27].
On the other hand, GaN transistors have even lower switching losses than SiC
devices but most products at the moment are rated for voltages up to 600 V,
which makes them counterparts of Si MOSFETs. Thus, GaN transistors are more
suitable for dc/dc converters and other lower-voltage-rated devices, where high
switching frequency is preferable [6, 34].

5
Chapter 1. Introduction

The main disadvantage of SiC power switches is their higher cost compared to
IGBTs, but this cost difference tends to reduce as SiC technology becomes more
widely used. The faster switching transients of WBG devices have as an additional
side-effect much higher dv/dt on their ac PWM voltage. This can impose an
additional stress on the insulation of the stator windings of the electrical machine,
in case the machine is driven by a SiC inverter [35]. Similar issue is observed at
the dc-link capacitors that need to filter current and voltage ripple with higher
transients.
Considerable research work has been conducted on the design of SiC inverters
suitable for motor drives, such as in [36]. However, there has not been any
comprehensive study about the design process of automotive SiC inverters with
high power density, where specific guidelines regarding the design of the whole
drive system are given. The design specifications of the vehicle should also be
considered and correct sizing of the inverter and its cooling circuit needs to be
properly implemented. Lastly, experimental testing of the drive in laboratory
conditions is necessary to validate the design.

1.1.2 Kite-based Tidal Power Generation System


Offshore power plants producing energy from the sea tides have attracted significant
research interest in the recent years, since this can be a renewable energy source
with large potential [37–40] and an alternative to the wind and solar energy sources.
Specifically, the United Kingdom and France cover almost 90% of the European
tidal resource with 48 GW of power, as reported in [37, 38]. Tidal energy has
considerable advantages compared to the other renewable energy sources, such
as the fully predictable nature of the tidal movements and its robustness against
climatic changes, such as the wind, rain, snow, clouds and fog [41].
Up to now, tidal power has been largely unexploited, although attempts to harvest
electric energy from sea tides date back to 1966, when the oldest tidal power plant
was installed in France [42]. Most of the traditional tidal power plants required the
construction of large dams and static undersea constructions, which had a large
investment cost as well as a high environmental impact for the local communities.
These are some of the reasons that have hindered until now further development
of tidal power projects in large scale.
An emerging hydrokinetic energy technology named "Tethered UnderSea Kite"
(TUSK) has been proposed as a new more efficient way to produce electrical energy
from the sea tides, using axial-flow water turbines mounted on a kite which can

6
1.1. Background

Fig. 1.2: TUSK plant schematic into the sea showing the tether connecting the kite to
the sea bed [Property of Minesto AB]

move freely inside the sea following the motions of the sea currents [43]. This
type of technology can increase the power production of a specific tidal turbine,
when compared to a traditional static tidal turbines of the same size. Therefore,
the potential resources of tidal power can be increased by using the TUSK-based
systems. Also, the environmental impact of the plant’s installation is limited. A
recent application of the TUSK system is the Deep Green [44, 45] concept which
has been developed by Minesto AB and is under development in various projects,
such as the PowerKite project [46]. The undersea kites can be installed either
standalone, as shown in Fig. 1.2, or they can build arrays.
The concept of using a moving kite for producing power has initially been proposed
for airborne systems that would increase the power density of existing wind power
turbines, as cited in [47–53]. As described there, the turbine is mounted on the kite,
which is attached to the ground through a tether. While the kite is travelling on
the air (in the case of airborne kites) or in the sea (for the undersea kites) following
a predefined trajectory, the relative speed of the air or water flowing through the
turbine becomes 5 to 8 times higher compared to a traditional static turbine. The
kite is moving fast following a trajectory that resembles to the "∞" symbol. Since
the power generated from the turbine is proportional to the third-power of the
relative wind/water speed, the electric power produced by the generator is highly
increased. Thus, the power density of the whole power generation system is also
increased.
The increase of the generated power density is a distinctive advantage for TUSK
systems compared to fixed marine turbine technologies [43] and this is one of the

7
Chapter 1. Introduction

characteristics that would make tidal power competitive against other renewable
energy sources. Additionally, the increased relative speed of the fluid flowing
through the plant’s turbine, allows TUSK-based tidal plants to operate at places
with medium or low tidal streams, which may not be cost-effective using the
traditional fixed tidal turbines that require high-speed tidal streams [44].

Optimal Control of Tidal Power Generator

The TUSK-based plants, however, have some special requirements for the design
and control of the power generation subsystem, which are going to be further
analyzed in this thesis. Firstly, the speed of the water flowing through the kite’s
turbine experiences fluctuations, which depend on the trajectory of the TUSK in
the sea. These speed fluctuations are periodic and cause variations in the output
electrical torque of the plant’s generator and of the produced power at the grid
side of the system [54]. The speed and power variations can be estimated using
analytical calculations, if the trajectory of the kite is known. A proper design of
the controllers for the power conversion system should be made in order to ensure
stability under this highly dynamic operation. Verifying the effectiveness of the
power converters’ control in laboratory conditions is necessary and, therefore, a
laboratory emulator of the kite’s power generation system is needed which will
represent the dynamics of the kite’s tidal turbine. Laboratory emulators of wind
power systems have been presented in [55–59], where the wind turbine has been
emulated by a motor and the control of the real wind power generator has been
tested. A similar experimental set-up can be used in this thesis to emulate the
operation of the TUSK-based tidal generator.
The Maximum Power Point Tracking (MPPT) algorithm of the kite’s tidal genera-
tor is another topic that has not been extensively investigated. Since the operating
conditions are similar to wind power plants, similar MPPT control strategies can
be applied for tidal power systems, as the ones investigated in [60–64]. However,
due to the highly dynamic operation of the kite and the constantly variating flow
speed of the water through its turbine, these methods should be studied and
their effectiveness should be experimentally evaluated. The turbine inertia is also
a factor that can influence the effectiveness of the MPPT control of the kite’s
generator. The inertia of the system creates an error between the actual rotational
speed of the generator and the reference speed, which is calculated by the MPPT
control. Therefore, improvements should be proposed to properly adjust the
generator speed control for the specific needs of the examined application.

8
1.1. Background

Medium Voltage Multilevel Machine Drives

An additional constraint of the TUSK systems is the limited available space inside
the nacelle of the kite for the generator and the power converters which makes
the design of the power conversion system a challenging task. Therefore, power
equipment of high power density is a requirement for this application, similarly
with the automotive drive described before. Also, the tether of the kite provides
both mechanical coupling of the system with the sea bed and electrical connection
to the grid. The weight and the diameter of the electrical cables inside the tether
should be kept as low as possible in order to reduce the drag that the tether creates,
when the kite is moving inside the sea. This makes the efficient transportation of
the produced active power from the kite to the grid challenging, since the power
of the system is high.
Increasing the operating voltage of the kite’s power generation system from
commonly used low voltage to the medium-voltage range can increase the efficiency
of the power transportation through the subsea cables, since this results to reduced
current and ohmic losses. Alternatively, the diameter of the subsea cables and
the tether of the kite can be reduced, due to the lower current requirements.
Medium voltage is defined by IEEE Std 1623-2004 [65] as the ac rms voltage
between 1-35 kV. The 2LC can still be used in medium voltage applications by
having high-voltage semiconductors or multiple series-connected low-voltage power
switches in each switch position. However, problems with the voltage balancing of
these switches during the switching transient are the main disadvantages of this
solution [66, 67]. Therefore, the use of multilevel VSCs is a preferable solution for
this case.
The main advantage of multilevel converters is the capability to use power semi-
conductors of low voltage ratings, which have lower cost and good commercial
availability [68, 69]. Another important advantage is the reduction of the harmonic
distortion at the point of common coupling (PCC) of the kite, which eliminates the
need for large passive filters at the grid-side of the system [68] that contribute a lot
to the total weight and size of the drive mounted inside the kite. An improvement
of the converter’s efficiency has also been reported in [70], when using multilevel
converters compared to 2LC of the same operating voltage and power. Multilevel
converters can also reduce the voltage derivative dv/dt, which is caused by the
switching transients of the converters, as stated in [71]. This decreases significantly
the produced EMI that could normally cause lots of operating problems on the
low-voltage electronics of the system.

9
Chapter 1. Introduction

The available multilevel topologies need to be compared with each other in order to
select the most suitable converter for the power conversion system of the undersea
kite. A comparative study based on literature review for multilevel converters
suitable for drive applications is conducted in Section 4.2 of the thesis and the
5-level Neutral Point Clamped (NPC) converter is finally chosen. The 5-level
NPC converter can increase the operating voltage of the drive system up to 4
kV (peak), in case power switches with voltage rating of 1.2 kV are used. The
5-level converter has four series-connected capacitor banks at the dc side and a
well known problem of this topology is the imbalance of the individual capacitor
voltages. This imbalance would create overvoltage in some of the power switches
of the converter and low frequency harmonics at the ac side of the system. Special
control and/or hardware is required to balance the capacitor voltages, which may
increase the complexity and cost of the system. Thus, several methods have been
proposed to alleviate this drawback and they can be divided into two types, the
software and the hardware based techniques.
The software-based balancing method is based on the modification of the standard
space vector modulator of the NPC converter, as suggested in [72–80]. However,
most of the solutions proposed in these studies require some limitations in the
operation of the drive, such as the difficulty to operate under unity power factor,
which is usually the preferred operation mode for the grid-side converter of the
drive. Some studies propose also the operation of the 5-level converter in a quasi-
2-level [81] and quasi-3-level mode [82] only in order to avoid specific switching
combinations that create unbalances of the dc-link voltages. However, this method
may increase the harmonics at the ac side of the NPC converters and the need for
grid filters, which makes it less suitable for several generator drives, such as the
one designed for the TUSK where the available space for the drive is limited.
Another option for balancing the capacitor voltages of the dc-link is by employ-
ing dc/dc converters that can actively distribute the voltage across the four
series-connected capacitors, as proposed in [83–88]. This technique has several
advantages; however, the cost and the losses of the extra power devices for the
dc/dc converters are serious drawbacks. From the above it is concluded that
further investigation should be conducted to develop a new technique for the
capacitor voltage balancing in a drive with 5-level NPC converters without the
previously mentioned disadvantages.

10
1.2. Aims of the Thesis

Wide-Bandgap Technology in Offshore Power Generation

WBG semiconductors and specifically SiC MOSFETs have been proposed for power
converters of future offshore power plants [89], since high-voltage capable SiC power
modules could help to increase the voltage rating of the power conversion system.
High-voltage SiC devices in 2LC topologies are also studied in [90] for active power
filters in offshore applications. However, as stated in [91], currently available
SiC power modules do not have large enough power capabilities to fulfil all the
requirements of large offshore power plants. Therefore, multilevel converters are
still the only solution to build systems with high voltage and high power. Further
development of the SiC technology in the future, with the commercialization of
high-voltage and high-power modules, might mean that converters for offshore
applications could be built using standard 2-level structure.
On the other hand, a power generation system that combines NPC-based machine
drive with the use of SiC power switches could benefit from the advantages of both
the multilevel converters (lower THD and utilization of low-cost low-voltage-rated
switches) and the WBG devices (low inverter losses). A reduction of the inverter
losses is highly beneficial for the TUSK-based tidal plant, since this would allow
the construction of more compact converters with less cooling requirements. This
concept will be investigated in this thesis.

1.2 Aims of the Thesis


The aim of this thesis is to investigate different design approaches for electrical
machine drives with SiC power inverters, which are suitable for applications where
high power density and efficiency are required.
Firstly, a medium voltage PMSM drive with 5-level NPC converters is designed,
which is intended to be used in kite-based tidal power generation systems. One
purpose of the thesis is to present suitable control for the converters, considering
the highly dynamic operation of the kite due to the periodic variations in the
active power and rotor speed. Another target of this work is to develop voltage
balancing techniques for the dc-link capacitors of the NPC converters, so that
their individual voltage can be effectively controlled.
The second aim of the thesis is to develop a comprehensive design approach for
800 V SiC automotive inverters. High power density in this case is also required,
which is achieved through an advanced design of the inverter’s mechanical structure,
such as the dc bus and the cooling system.

11
Chapter 1. Introduction

A final goal of the project is to verify the research outcomes through simulations
of each system, as well as through experimental tests on laboratory prototypes.

1.3 Contributions
The key scientific contributions of the thesis can be identified as follows:
• MPPT generator control for TUSK power plants:
A properly designed MPPT controller for the generator of a TUSK tidal
power plant has been implemented that can follow effectively the power and
speed variations. The developed controller has been tested experimentally on
a 35 kVA laboratory set-up where the operation of the turbine is emulated
by an induction motor (IM).
• DC-link voltage balancing algorithms for 5-level NPC converter drive:
New voltage control techniques have been proposed that can balance the
capacitor voltages on the dc-link of a generator drive with 5-level NPC con-
verters. An optimized selection of the redundant states on the SVM-plane
has been utilized as the basic tool to regulate the capacitor voltages. This is
combined with hardware-based balancing using extra dc/dc converters in
order to ensure effective voltage control under every operating condition of
the drive, without any limitation in the operation of the converters. The de-
veloped control requires less components and can reduce the losses compared
to conventional techniques that utilize only hardware-based balancing.
• Medium voltage PCB design for NPC inverter with SiC switches:
A generator drive with two back-to-back (B2B) connected 5-level NPC
converters has been developed, which can be used at the power conversion of
a TUSK plant. A laboratory prototype of the generator drive with maximum
designed dc voltage up to 4 kV has been firstly simulated and then designed
and built during this project. The system is tested experimentally with
dc-link voltage up to 670 V.
The power converter PCBs, measurement devices and control platforms of
the prototype system have been specially designed for low EMI and ringing
on the power switches. Also, SiC MOSFETs and diodes have been used in the
NPC converters, which allows further reduction of the losses in the system.
Suitable gate driver boards for the SiC MOSFETs have been designed that
achieve high voltage insulation and low EMI on the control signals.
• Development of high power 49 kW/liter SiC automotive inverter:

12
1.4. Thesis Outline

The design procedure of a 450 A SiC automotive inverter has been studied
and a new iterative design approach for the liquid cooling components of
the inverter has been developed. Transient CFD computations are utilized
in combination with simulations of legislative testing procedures for electric
vehicles, such as acceleration tests and driving cycle analysis, in order to
fine-tune the inverter based on the needs of the specific vehicle. A special
design of the dc-link capacitor bank has also been developed in order to
further reduce the volume of the system.
• Zero-sequence control strategy for 2LC-based P-HIL systems:
A machine emulation testing platform for PMSM drives with reduced com-
ponent count has been proposed. The electrical machine has been emulated
by an additional VSC, which is connected to the ac side of the main inverter
of the drive and shares the same dc-link. A new control scheme of the
common mode voltage of the resultant loop has been developed, so that the
zero-sequence current which normally flows in this circuit can be suppressed
without the use of hardware filters. Instead, the redundant states of the
SVM have been utilized to regulate this current.
A P-HIL testing platform has been built with the previously mentioned
450 A SiC inverter in order to experimentally verify the functionality of the
control.

1.4 Thesis Outline


The thesis is organized into eight chapters:
• Chapter 1 is the introduction of the thesis and describes the motivation, the
aim and the main contributions of this study. The background applications
of this PhD project, tidal power generation and electrified transportation,
are also described.
• Chapter 2 provides an overview of the vector control strategies used in
the remaining part of the thesis. The vector control structure for grid-
connected inverters and field-oriented-control for synchronous machines are
also analyzed.
• In Chapter 3 an overview of the first background application of this project,
the TUSK tidal generator, is presented. The tidal power generation system
has been modelled with analytical equations, which are described in detail.
The control system for both converters of the grid-connected drive is also

13
Chapter 1. Introduction

designed and experimental results from a low-scaled tidal power emulator


are shown.
• Chapter 4 starts with an analysis of the most well-known multilevel converter
topologies for drive applications. Modulation strategies for N -level multilevel
converters are presented. Capacitor voltage balancing strategies for machine
drives with 5-level NPC converters are presented utilizing the Space Vector
Modulation.
• Experimental verification of some of these voltage balancing control tech-
niques is presented in Chapter 5, where the design process of a medium
voltage prototype multilevel drive is analyzed.
• Machine emulators for testing electric drives are studied in Chapter 6. Space
Vector Modulation is utilized to control the common mode voltage of the
topology and eliminate the circulating zero-sequence current. Experimental
results on a power hardware-in-the-loop set-up emulating a 60 kW PMSM
are presented.
• Chapter 7 presents the second background application of this project, the
design of high-power inverters with high power density for electric vehicles
using SiC power switches. The complete design approach of the inverter is
presented, including electrical and thermal modelling by utilizing FEM and
CFD simulation tools. The machine drive of a test-case vehicle has also been
modelled in order to fine-tune the inverter design to meet the requirements
of the specific application.
• Finally, the conclusions of the thesis and suggestions for future work are
reported in Chapter 8.

1.5 List of Publications


The peer-reviewed publications originating from this thesis are:
I. G. Mademlis, Y. Liu, and N. Saadat, "Combined voltage balancing tech-
niques of the DC link in five-level medium voltage NPC back-to-back convert-
ers for offshore renewable generation," in Proc. 19th European Conference
on Power Electronics and Applications (EPE’17 ECCE Europe), 2017, pp.
P.1-P.10, doi: 10.23919/EPE17ECCEEurope.2017.8099317.
II. G. Mademlis, Y. Liu, and J. Zhao, "Comparative Study of the Torque
Ripple and Iron Losses of a Permanent Magnet Synchronous Generator

14
1.5. List of Publications

Driven by Multilevel Converters," in Proc. 23rd International Conference


on Electrical Machines (ICEM), September 3-6, 2018, pp. 1406-1412, doi:
10.1109/ICELMACH.2018.8506775.
III. G. Mademlis, Y. Liu, P. Chen, and E. Singhroy, "Generator Speed Control
and Experimental Verification of Tidal Undersea Kite Systems," in Proc.
23rd International Conference on Electrical Machines (ICEM), September
3-6, 2018, pp. 1531-1537, doi: 10.1109/ECCE.2018.8558080.
IV. G. Mademlis and Y. Liu, "DC Link Voltage Balancing Technique Utilizing
Space Vector Control in SiC-based Five-Level Back-to-Back-Connected NPC
Converters," in Proc. of the IEEE Energy Conversion Congress and Exposi-
tion (ECCE), September 23-27, 2018, pp. 3032-3037,
doi: 10.1109/ECCE.2018.8558080.
V. G. Mademlis and Y. Liu, "Feed-forward Control of Active Voltage Balancing
Converter in Electric Drive with Five-Level NPC Converters," in Proc. of the
2018 IEEE Energy Conversion Congress and Exposition (ECCE), September
23-27, 2018, pp. 7258-7264, doi: 10.1109/ECCE.2018.8557365.
VI. C. Larsson, A. Rydgård, G. Mademlis, Y. Liu, and M. Fredriksson, "Over-
voltage Mitigation of Medium Voltage Electric Drives with Long Cables using
Multilevel-Converters and Passive Filters," in Proc. of the 21st European
Conference on Power Electronics and Applications (EPE’19 ECCE Europe),
2019, P.1-P.10, doi: 10.23919/EPE.2019.8915441.
VII. G. Mademlis, Y. Liu, P. Chen, and E. Singhroy, "Design of Maximum
Power Point Tracking for Dynamic Power Response of Tidal Undersea Kite
Systems," in IEEE Trans. Industry Applications, vol. 56, no. 2, pp. 2048-
2060, March-April 2020, doi: 10.1109/TIA.2020.2966189.
VIII. N. Sharma, Y. Liu, G. Mademlis, and X. Huang, "Design of a Power
Hardware-in-the-Loop Test Bench for a Traction Permanent Magnet Syn-
chronous Machine Drive," in Proc. 24th International Conference on Electri-
cal Machines (ICEM), August 23-26, 2020, pp. 1765-1771,
doi: 10.1109/ICEM49940.2020.9270838.
IX. G. Mademlis, Y. Liu, J. Tang, L. Boscaglia, and N. Sharma, "Performance
Evaluation of Electrically Excited Synchronous Machine compared to PMSM
for High-Power Traction Drives," in Proc. 24th International Conference
on Electrical Machines (ICEM), August 23-26, 2020, pp. 1793-1799, doi:
10.1109/ICEM49940.2020.9270852.

15
Chapter 1. Introduction

X. G. Mademlis, R. Orbay, Y. Liu, and N. Sharma, "Designing Thermally Uni-


form Heatsink with Rectangular Pins for High-Power Automotive SiC Invert-
ers," in Proc. IECON 2020 - 45th Annual Conference of the IEEE Industrial
Electronics Society, 2020, pp. 1317-1322, doi: 10.1109/IECON43393.2020.9254692.
XI. G. Mademlis, Y. Liu, N. Sharma, and X. Huang, "Circulating Current
Reduction in Common DC-Link Power-HIL for Drives using SVM with
Zero-Sequence Compensation," in Proc. IECON 2020 - 45th Annual Con-
ference of the IEEE Industrial Electronics Society, 2020, pp. 4673-4678, doi:
10.1109/IECON43393.2020.9254744.
XII. L. Boscaglia, N. Sharma, Y. Liu, and G. Mademlis, "Balancing Peak-
torque and Drive-cycle Efficiency with Magnet Dimensioning of Permanent
Magnet Synchronous Machines," in Proc. IECON 2020 - 45th Annual
Conference of the IEEE Industrial Electronics Society, 2020, pp. 883-888,
doi: 10.1109/IECON43393.2020.9255349.
XIII. G. Mademlis, R. Orbay, Y. Liu, N. Sharma, R. Arvidsson, and T. Thiringer,
"Multidisciplinary Cooling Design Tool for Electric Vehicle SiC Inverters
Utilizing Transient 3D-CFD Computations," in eTransportation journal,
Elsevier, vol. 7, pp. 100092, November 2020, doi:10.1016/j.etran.2020.100092.
XIV. G. Mademlis, N. Sharma, Y. Liu, and J. Tang, "Zero-Sequence Current
Reduction Technique for Electrical Machine Emulators with DC-Coupling by
Regulating the SVM Zero States," under review since Feb. 2021 in IEEE
Trans. Industrial Electronics.
XV. J. Tang, D. Pehrman, G. Mademlis and Y. Liu, "Common Mode Power
Control of Three-Phase Inverter for Auxiliary Load without Access to Neutral
Point," under review since Feb. 2021 in IEEE Trans. Power Electronics.

16
Chapter 2

Synchronous Machine Drives:


Topologies and Control

Electric drives with synchronous machines are studied in this thesis and vector
control is used for the three-phase voltage source converters (VSC). The definition
and the basic structure of the drive system are presented in this chapter, when it
is connected to the grid and when it is connected to a dc source. Vector control
theory for the converters in both cases is presented.

2.1 Structure of Electrical Machine Drives


An electrical machine is a device that converts mechanical to electrical energy.
Depending on the direction of energy flow, the electrical machine can operate as
a generator (producing electrical energy) or as a motor (producing mechanical
energy). A three-phase ac machine needs an ac voltage source connected to the
three terminals of its stator windings. Since this ac voltage source does not always
have the correct voltage amplitude and frequency, a power converter is required
to provide that. The power converter is composed of switching devices and acts
as an intermediate step between the machine and the electrical source. It is
responsible for processing and controlling the flow of electrical energy by supplying
the machine with the correct voltage. Bidirectional power flow is necessary for
the converter, since the electrical machine can operate both as a motor and as a
generator.
An electric drive system comprises of an electrical machine and a power converter,
as the schematic layout shown in Fig. 2.1(a). The right side of the power conversion

17
Chapter 2. Synchronous Machine Drives: Topologies and Control

power flow
mechanical
load ac
M or +
-
power
converter source
(a)
power power
(+) Vdc tot (+) Grid (+)
Tm ie abc is abc Tm ie abc Battery
M Zf M +

vs abc -

(b) (c)

Fig. 2.1: Simple schematic diagram of (a) an electrical machine connected to a power
converter, (b) a grid-connected machine drive with two back-to-back connected
converters (typical for power generation systems) and (c) machine drive supplied
by a dc battery source (typical for electric vehicles)

system is connected to a power source, which can have either dc or ac voltage,


depending on the application.
In case the drive system is connected to the grid, the power conversion system is
composed of two dc/ac converters, which are connected back-to-back. The first
converter at the grid-side rectifies the ac voltage of the grid to an intermediate
dc-link and the second converter is responsible for controlling the electrical machine.
When only single-direction of power is needed, which flows from the grid towards
the machine operating only as a motor, a passive diode rectifier is enough as the
grid-side converter. However, when bidirectional power flow is required, such as
in power generation systems, an active-front-end inverter is used at the grid-side
in order to synchronize it to the grid and control the power exchange [see Fig.
2.1(b)].
In applications where a dc power source is connected directly to the dc-side of the
machine converter as in Fig. 2.1(c), only the left-side converter of Fig. 2.1(b) is
utilized. A representative example is the drive systems of electric vehicles, where
the machine drive is supplied by a battery. The battery model has not been
studied in this project and it is considered as an ideal dc voltage source.
The energy flow marked with blue arrow in Figs. 2.1(b)-(c), with direction exiting
the converters at their ac-side, is considered positive for the following analysis.

18
2.2. Field Oriented Control

2.2 Field Oriented Control


The Field Oriented Control (FOC) method for synchronous machines consists
of two cascaded control loops, with the outer loop being the speed control of
the machine and the inner loop the current control. In order to implement the
following control strategy, measurements of the machine stator current ie , rotor
angle θr and dc-link voltage Vdc tot are needed.

2.2.1 Speed Controller Design


A schematic of the machine speed controller is shown in Fig. 2.2. It consists of
a proportional-integral (PI) control loop and an active damping term Baω . The
gains of the PI loop are chosen based on the Loop Shaping Method [92] in order
to have smooth response of the controller, similar to a first-order system. They
are defined as

KP ω = aω J (2.1)
KIω = aω (b + Baω ) (2.2)

where aω is the closed-loop bandwidth of the speed controller, J is the inertia and
b is the friction constant of the machine drivetrain. The active damping term Baω
is selected as
Baω = aω J − b (2.3)

in order to increase the robustness of the controller against estimation errors of


the system parameters.
The output of the speed controller is the reference electromagnetic torque for the
machine Te∗ which is limited up to its rated value by a saturation block. Anti-reset
windup is also implemented, which is not shown in the block diagram of Fig. 2.2.
The speed response of the machine is described by its mechanical model, when

Speed Controller ie d ie q

 *
r
KI ω 1/s Current ref.
+ + Saturation look-up-table Machine 3-phase
ωr - ve*
θr PLL KP ω + Current SVM machine
- Te* ie* dq Controller VSC
Βα ω ωr

Fig. 2.2: Block diagram of speed control loop

19
Chapter 2. Synchronous Machine Drives: Topologies and Control

the electromagnetic torque Te and load torque TL are applied.


dωr T ∗ ≈Te dωr
J + b ωr + T0 = Te − TL ==e==⇒ Te∗ = J + b ωr + TL (2.4)
dt T0 =0 dt
where T0 is the static friction constant measured in Nm and can help to increase
the fidelity of the machine mechanical model.
The rotational speed of the machine ωr can be calculated through derivation of the
rotor angle ωr = dθr /dt. However, a speed estimator based on the phase-locked-
loop (PLL) principle, as described in [93], can offer more stable calculation of the
machine speed and is preferred in this study.

2.2.2 Current Reference Generation


The electromagnetic torque Te produced by a synchronous machine can be calcu-
lated as
3
Te = p(ψd ieq − ψq ied ) (2.5)
2
where p is the number of pole-pairs. If the inductances of the machine are
considered constant, the torque equation becomes
3
Te = P [ψmd ieq + (Ld − Lq ) ied ieq ] (2.6)
2
The first term of the above equation depends on the q-current and the magnetizing
flux linkage from the rotor ψmd and can be named synchronous torque. The second
term of (2.6) depends on the difference between the d- and q-axis inductances and
is named reluctance torque.
Non-salient synchronous machines, such as the surface-mounted PMSM, have equal
d- and q-axis inductances Ld = Lq and, therefore, they have only synchronous
torque. For salient permanent magnet machines, such as for the IPMSM, the
inductances are Ld < Lq and the total machine torque is the sum of reluctance
and synchronous torque.
Many methods exist for selecting the optimal combination of dq-axis stator currents
in order to create the requested torque of the machine. The current and voltage
operating limits of the system need also to be considered, when defining the current
reference [94]. In case the synchronous machine has controllable excitation with
rotor field current if , there is an additional degree of freedom for controlling the
machine [14]. Optimal selection of these variables allows the minimization of the
machine losses.

20
2.2. Field Oriented Control

A simple way to calculate the required reference current is by setting i∗ed = 0 A and
controlling the machine just with the synchronous torque. This gives the optimal
current reference of non-salient machines, since the term Te /ie is maximized and
the drive operates at MTPA with minimum stator copper losses.
However, a non-zero d-current needs to be applied for salient machines, in order
to utilize also the reluctance torque [95, 96]. MTPA operation can then be
implemented in salient machines. The current reference generation methods can
be classified into three general categories:
1. The current reference is generated using analytical equations for minimizing
the machine losses, as in [97]. These equations are based on the machine
parameters and the current reference is calculated online, while the drive
operates.
2. A search algorithm calculates offline the optimal dq-current combination for a
finite number of torque-speed points on the machine operating map utilizing
a detailed machine model. FEM computation results of the machine losses,
inductances or flux linkages and inverter losses are used. The calculated
current is stored in a look-up-table which is loaded on the control board
of the inverter. Then, the correct reference is obtained by the MCU with
interpolation in this look-up-table, where the inputs are the machine reference
torque and speed. The temperature can also be considered as an input
parameter of the table, as shown in [98].
3. MTPA point tracking is implemented by injecting an additional high-
frequency signal, as presented in [96, 99, 100].
The first two methods have been used for the electrical machine control of this
thesis. Analytical current reference generation is applied in the medium voltage
PMSM drive of Chapter 5, while look-up-tables of the optimal current reference
are used in Chapters 6-7.

2.2.3 Machine Current Controller Design


The terminal voltage of the synchronous machine can be expressed as follows
died died
ved = Rs ied + Ld − ωe ψq ⇒ ved = Rs ied + Ld − ωe Lq ieq (2.7)
dt dt
dieq dieq
veq = Rs ieq + Lq + ωe ψd ⇒ veq = Rs ieq + Lq + ωe Ld ied + Ef (2.8)
dt dt
where Ld and Lq are the no-load d- and q-axis stator inductances, Rs the per-
phase stator resistance, ωe the electrical frequency in rad/s and Ef the back-EMF

21
Chapter 2. Synchronous Machine Drives: Topologies and Control

d-axis current controller

KI d 1/s
+
* + + ee* d + ve* d +
i ed KP d
- - -
ie d Βα d
ie q Lq x ve*
ieq·Lq·ωe ve* dq dq 3-phase
ωe
ωr p Vdc tot αβ SVM machine
θe VSC
ied·Ld·ωe
ie d Ld x
θe
ie q Βα q electrical
- + - angle
+ + ee* q ve* q +
ie* q KP q
+
+ +
KI q 1/s
Ef EMF

q-axis current controller

Fig. 2.3: Block diagram of electrical machine current controller in dq-frame

voltage.
The current control is the inner control loop of the machine-side converter and
calculates the reference voltage for the machine based on the above equations
(2.7)-(2.8). The block diagram of the controller is shown in Fig. 2.3. The current
controller split between the d- and q-axis control blocks. Inputs are the current
references i∗edq and actual current values iedq in both d- and q-axis. The gains and
the active damping term of the d-axis PI controller are selected based on the Loop
Shaping Method as follows

KP d = Ld ac (2.9)
KId = ac (Rs + Bad ) (2.10)
Bad = Ld ac − Rs (2.11)

where ac the bandwidth of the controller. Similar equations define the parameters
of the q-axis PI controller replacing in (2.9)-(2.11) the Ld inductance with Lq .
Common practice is to select the bandwidth of the current controller at least an
order of magnitude lower than the switching frequency, i.e. ac = 2π f10
sw
, in order
to have a stable system during dynamic and steady-state operation.
A feed-forward of the calculated machine voltage is added to the output of the

22
2.3. Vector Control for Grid-Side Converter

PI-regulators e∗ed and e∗eq according to the following formulas

v∗ed = e∗ed − Lq ieq ωe (2.12)


v∗eq = e∗eq + Ld ied ωe + Ef (2.13)

In case the synchronous machine has permanent magnets, the back EMF is
calculated by
Ef P M = ψmd ωe (2.14)

where ψmd is the magnet flux linkage. In case we have a synchronous machine
with rotor windings
Ef = Lmd if ωe (2.15)

where Lmd is the d-axis magnetizing inductance and if the rotor field current.
The reference voltage v∗e dq , which is the output of the current control loop, is
limited by a saturation block in order to avoid overmodulation. This limits the
voltage amplitude, but does not modify the voltage angle. Anti-reset windup is
also implemented, which is not shown in the block diagram of Fig. 2.3.
Space vector modulation is finally used for generating the gate signals of the power
switches in the converter.

2.3 Vector Control for Grid-Side Converter


Vector control in the dq-frame is used for the grid-side converter. Two cascaded
control loops exist, with the outer loop regulating the voltage of the dc-link
capacitor bank and the inner loop regulating the d- and q-axis currents. The
converter is synchronized with the grid through a PLL block.
A three-phase passive filter consisting of an inductor Lf with its series resistance
Rf is considered to be installed between the ac output of the converter and the
PCC.

2.3.1 Real-/Reactive-Power Control Design


The exchange of power with the PCC of the grid can be described as follows
3
Pgrid = [vsd isd + vsq isq ] (2.16)
2
3
Qgrid = [−vsd isq + vsq isd ] (2.17)
2
where vsd and vsq are the d- and q-axis measured voltages of the grid at the PCC.

23
Chapter 2. Synchronous Machine Drives: Topologies and Control

The converter is synchronized with the grid voltage through the PLL shown in Fig.
2.4(a), which tracks the grid angle θe grid and frequency ωe grid . The PLL consists
of a PI controller that aligns the grid voltage vector with the d-axis. Therefore, if
the controller is in steady state, it can be assumed that vsq=0 . The gains of the
PLL controller are selected as described in [93, 101]

KP P LL = 2 aP LL (2.18)
KIP LL = a2P LL (2.19)

where aP LL is the bandwidth of the PLL controller.


By considering vsq = 0, the previous power equations in (2.16)-(2.17) become
3
Pgrid = vsd isd (2.20)
2
3
Qgrid = − vsd isq (2.21)
2
Therefore, it can be seen that the active power injected by the converter to the grid
Pgrid can be controlled by the d-axis current, while the q-axis current regulates
the reactive power Qgrid .
Since the grid-side converter is used in back-to-back connection with another
three-phase converter controlling the electrical machine of the drive, the power
balance at the dc-link of the topology can be expressed as follows [102], based on
the energy-flow direction labeled in Fig. 2.1(b) with blue arrows,
2
Cdc tot dVdc tot
= −Pmachine − Pgrid − Ploss (2.22)
2 dt

angle of measured KI PLL 1/s ωe grid


grid voltage
tan 1 Vs2d  Vs2q  KP PLL + 1/s θe grid
dθe grid
frequency Limit within
limiter [-π, π]

(a)
KI dc 1/s Limit within
Vdc2 tot + P* ± Is max
Vdc tot u2 KP dc + grid
2/3
x
is*d
+ ÷
-
x
*
Qgrid - 2/3 is*q
Vdc* tot u2
÷
vs d

(b)

Fig. 2.4: Block diagram of the grid-side VSC (a) PLL and (b) dc-link controller

24
2.3. Vector Control for Grid-Side Converter

where Pmachine the power of the machine-side converter and Ploss the power losses
of the system. The left part of the above equation corresponds to the rate of
change in the energy of the dc-bus capacitors. Since both the Pmachine and the
Ploss cannot be directly controlled by the grid converter, it is clear that the dc-link
voltage dynamics Vdc2
tot depend on the Pgrid , which can be regulated by the isd
as seen in (2.20).
The dc-link controller, shown in Fig. 2.4(b), keeps the dc-link voltage close to its
reference value Vdc

tot . The proportional gain of the controller is selected as

KP dc = aDClink · Cdc tot (2.23)

where aDClink is the controller bandwidth and is usually selected an order of


magnitude lower than the current controller bandwidth, described in the following
subsection. In this project, the dc-link PI-controller is selected to have dynamic
response with time constant of 0.03 sec/rad and, therefore, the integrator is

KP dc
KI dc = (2.24)
0.03
Output of the dc-link control loop is the reference current of the grid converter.

2.3.2 Grid-side Current Controller Design


The grid voltage at the PCC can be expressed, transformed in the dq reference
frame, as follows
disd
vsd = vtd − Rf isd + Lf ωe grid isq − Lf (2.25)
dt
disq
vsq = vtq − Rf isq − Lf ωe grid isd − Lf (2.26)
dt
where Lf and Rf are the inductance and resistance of the three-phase grid filter,
respectively, vtd and vtq are the d- and q-axis voltages at the ac side of the grid-side
converter.
The reference voltage of the converter is calculated by the grid-side current
controller and its block diagram is shown in Fig. 2.5. The proportional and
integral gains of the PI control loops for both d- and q-axis are KP s and KIs . The
current dynamics for both axes are identical and are described with the following
loop-gain transfer function [102]

KP s s + KIs 1
Hs (s) = HsP I (s) · Hf (s) = · (2.27)
s Lf s + Rf

25
Chapter 2. Synchronous Machine Drives: Topologies and Control

where HsP I (s) describes the PI control loop and Hf (s) the dynamics of the grid-
filter. The gains are selected based on the Loop Shaping Method, so that the
resultant closed-loop current dynamics of the converter have first-order response.

KP s d = KP s q = Lf acs (2.28)
KIs d = KIs q = acs (Rf + Bas ) (2.29)
Bas = Lf acs − Rf (2.30)

where acs is the controller bandwidth and Bas the current controller active damping.
Common practice is to select the current controller bandwidth at least an order of
magnitude lower than the switching frequency of the grid converter.
Feed-forward of the measured grid voltage vs for both axis is used, after being
filtered by an LPF. Feed-forward of the voltage drop across the grid filter inductance
is also applied, based on the voltage equations of (2.25)-(2.26)

VLf = (Lf ωe grid )j · is


(2.31)
= (−Lf ωe grid isq ) + (Lf ωe grid isd )j

The reference voltage of the grid converter v∗s dq is the output of the current
control loop, which is limited by a saturation block. The saturation block limits
the voltage amplitude, but does not modify the voltage angle. Anti-reset windup
is also implemented, which is not shown in the block diagram of Fig. 2.5.
Finally, the space vector modulation block generates the gate signals for the power
switches of the grid converter.
d-axis current controller vs d

LPF
(from dc-link KIs d 1/s
controller) + +
+ + +
vs* d +
is* d KPs d
- - -
is d Βα g
is q Lf x limiter vs*
isq·Lf·ωe grid vs*dq dq 3-phase
(from PLL) ωe grid Vdc tot αβ SVM Grid-side
isd·Lf·ωe grid θe grid VSC
is d Lf x
θe grid
is q Βα g
(from grid
- + - PLL)
+ + + vs* q +
is* q KPs q
+
+
KIs q 1/s LPF

q-axis current controller vs q

Fig. 2.5: Block diagram of the grid-side VSC current controller

26
Chapter 3

Modelling and Control of Kite-based


Tidal Power Generator Systems

The inverter control of a generator drive for kite-based tidal power systems is
analyzed in this chapter. Firstly, the motion of the kite and the power generation
profile are modelled. Afterwards, the control scheme of the inverters is described
and tested on a low-scaled laboratory emulator of the real kite.

3.1 Introduction to Control Techniques of the Kite


The power conversion system of the kite is mounted inside the nacelle and consists
of a generator, two B2B connected VSCs and a passive LCL filter at the point
of connection to the grid. A synchronous generator (SG) has been used for the
purpose of this study. A block diagram of the kite can be seen in Fig. 3.1(a). The
MPPT algorithm of the TUSK is an essential part of the generator control system.
The speed of the water flowing though the turbine is constantly changing, because
of the kite’s crosscurrent motion into the sea. The optimum reference generator
speed should be evaluated in order to extract the maximum electric power from
the water.
The MPPT algorithms presented in [60, 61] have as input the calculated active
power at the terminals of a permanent magnet synchronous generator (PMSG)
and calculates the reference speed of the rotor which serves as input of the
generator’s speed controller. Loss minimization control of the generator through
optimal current reference generation (see Subsection 2.2.2) can also be applied

27
Chapter 3. Modelling and Control of Kite-based Tidal Power Generator Systems

simultaneously. Another type of MPPT algorithm is described in [62–64], which


receives as input the measured rotational speed of the generator and generates the
optimal reference electrical torque that the current controller of the machine needs
to create. The speed of the rotor is controlled indirectly here, since the rotational
speed and the mechanical torque of the rotor will settle down to an equilibrium
with an optimal torque point, where the turbine operates at the maximum power
point. The speed control strategy of the kite’s generator is designed in this chapter
based on these two MPPT control schemes and their performance is compared
experimentally.
The rotational inertia of the tidal power system may affect the accuracy of the
applied MPPT control, since the generator speed controller gains are selected
based on the inertia of the mechanical drivetrain. This characteristic has a positive
influence on the wind turbine drives, where the wind fluctuations are stochastic
and cannot be predicted. The large wind turbine inertia has been utilized to
smooth out the power fluctuations caused by sudden changes of the wind speed
[103, 104] or to filter the highly fluctuating power output of wave generators [105].
The TUSK tidal system experiences also power fluctuations; however, contrarily
to the wind and wave power plants these power fluctuations are periodic, highly
predictable and dependent mainly on the predefined movement of the kite in
the sea. It is possible to smooth out the power fluctuations of a TUSK farm by
combining the power output of multiple kites, as shown in the Master thesis of
[54]. Therefore, it is not necessary anymore to utilize the turbine shaft inertia in
order to limit these power variations of the individual kites and the maximum
power should be extracted by the tidal generator at any time instant. The inertia
of the TUSK turbine shaft will still affect the transient response of the generator
speed controller, which could hinder it from following the optimal MPPT curve.
A speed control strategy needs to be applied on the TUSK tidal power generator
that has fast dynamic performance, so that the turbine of the kite can follow the
MPPT curve with the least possible error. A further analysis of the influence of
the shaft inertia is conducted at the experimental results of this chapter.

3.2 Tidal Power Plant System Modeling


The TUSK plant is controlled to undergo periodic movements into the water and
the orbit is eight-shaped. Water flows through the turbine blades of the kite
and creates a rotational torque that generates electric power at the output of
the generator. The generator is attached to back-to-back-connected VSCs which

28
3.2. Tidal Power Plant System Modeling

TUSK Plant

Va
DC-link Filter
Gearbox PCC Grid
3-phase 3-phase L1 L2
Tidal Gen-side Grid-side
Turbine SG VSC VSC
Cf

nturb i vgen Vdc tot is vs


θgen ωgen gen

(a)

(b)

Fig. 3.1: (a) Block diagram of the power conversion subsystem in a TUSK tidal plant;
(b) CAD-drawing of an early-concept kite for the DG500 project [106]: 1) tidal
turbine, 2) gearbox, 3) generator, 4) power conversion subsystem and grid
filters, 5) auxiliary equipment [Property of Minesto AB]

control the generator speed and the synchronization to the grid at the electrical
output of the topology. In order to calculate the produced electric power of the
topology, a mathematical model of the kite’s trajectory is needed.
A detailed analysis of the kite’s hydrodynamics can be found in the Master Thesis
report [54]. The model presented here is simplified and is utilized only for the
modelling of the electrical installation and the control of the power conversion
system. The exact geometry of the kite and its hydrodynamic performance have
not been modelled and the kite’s trajectory is considered to be unaffected by the
operation of the electrical generator. Also, the kite is considered to follow its
trajectory without error and the control mechanisms for that are not studied here.

29
Chapter 3. Modelling and Control of Kite-based Tidal Power Generator Systems

3.2.1 Modeling of Kite’s Trajectory and Power Generation


The three-dimensional depiction of the kite’s trajectory can be mathematically
described by the Viviani’s Curve, which is defined as the intersection between a
cylinder and a sphere, as shown in Fig. 3.2. The Viviani’s curve offers the minimum
curvature and maximizes power production compared to other trajectories, such as
the circle. The parametric equation of the Viviani’s Curve in the xyz-coordinate
system is given by

(a)

(b)

Fig. 3.2: Three-dimensional drawing of (a) the Viviani’s curve used as the trajectory of
the kite and (b) the TUSK plant following its trajectory P~r with the arrows
showing the direction of the kite’s movement after applying the rotation matrix
in (3.2)

30
3.2. Tidal Power Plant System Modeling

 
 x(t) = (R − r) + r · cos(t) 
  
~ (3.1)
p
P x(t), y(t), z(t) = y(t) = 2 · (R − r) · r · sin( 2t )
 
z(t) = r · sin(t)
 

where r is the cylinder radius, and R is the sphere radius. The parameter t varies
from 0 to 4π and some characteristic values of t on the Viviani’s curve can be seen
in Fig. 3.2(b). The x-axis of the graph is in-line with the tidal current flow.
The TUSK plant is elevated from the sea floor. Also, the tether of the TUSK and
the sea floor have an altitude angle D, as can be observed in Fig. 1.2. Therefore,
the position vector P~r can be rotated by multiplying with the following matrix

cos(D) 0 − sin(D)
 

P~r =  (3.2)
 ~
0 1 1 ·P

sin(D) 0 cos(D)

The partial derivative vector of the rotated position vector P~r can be defined as

∂ P~r ∂Pr ~ ∂Pr ~ ∂Pr ~


= i+ j+ k (3.3)
∂x∂y∂z ∂x ∂y ∂z
The kite’s trajectory is discretized into a definite number of points m and each
trajectory point is named with the variable i. A frame of reference is then defined
at each trajectory point, which is a local coordinate system describing the plant’s
direction with normalized matrix Bi that has dimensions {3 x 3 x m}.

Bi = [e~1i e~2i e~3i ] (3.4)

where e~1i , e~2i , e~2i are the normalized direction vectors of the three axes. These
three vectors point towards the forward direction, the wing axis and the tether
axis, respectively.
The partial derivatives of the position vector in (3.3) and the plant frame of
reference (3.4) are used to define the angular velocity matrix γ˙i , which describes
the rate of change of the kite’s pitch, roll and yaw on each point i of the trajectory

γ˙i = [∆P itchi ∆Rolli ∆Y awi ] (3.5)

A more detailed analysis of the mathematical movement equations can be found


in Chapter 5 of [54].
The mechanical power extracted by the tidal turbine can be calculated as follows
[104]
1
Pmi = ρAT Cp Va3i − Plossi (3.6)
2

31
Chapter 3. Modelling and Control of Kite-based Tidal Power Generator Systems

where ρ = 1025 kg/m3 is the salt water density, CP is the coefficient of performance
of the turbine, AT = πRT2 the swept area, RT is the turbine radius and Va the
effective linear water speed flowing through the turbine in m/s. The last term
Plossi describes the power losses at each trajectory point i owed to the motion of
the kite and can be calculated by

Plossi = Li T · γ˙i (3.7)

where Li is a torque coefficient matrix dependent on the kite geometry. The


following coefficients have been used for the analysis of this study and have been
empirically chosen
Li = [1500 2000 1000] (3.8)

The torque coefficients depend mainly on the design and the physical dimensions
of the whole tidal kite system and the empirical choice of these parameters could
introduce some error, which is not considered during this study. The tip speed
ratio T SRi of the tidal turbine at each trajectory point i is defined as
ωturbi RT
T SRi = (3.9)
Vai

where ωturbi is the turbine rotational speed in rad/s. By using (3.6) and (3.9), the
resultant power is
ρCP πRT5 3
Pm i = ω − Plossi (3.10)
2 · T SRi3 turbi
The active power at the point of connection to the grid can also be expressed as
follows
ρCP πRT5 3
Pgridi = ω − Plossi − Pe lossi (3.11)
2 · T SRi3 turbi
where Pe lossi is the power losses of the electric drive system (mechanical and
electrical) between the generator shaft and the connection to the grid. The turbine
torque Tmi and the electrical torque at the grid output of the kite Tgridi are
defined as
Pm i ρCP πRT5 2 Plossi
Tmi = = ω − (3.12)
ωturbi 2 · T SRi3 turbi ωturbi
Pgridi ρCP πRT5 2 Plossi Pe lossi
Tgridi = = 3 ωturbi − − (3.13)
ωturbi 2 · T SRi ωturbi ωturbi

The effective water speed Va depends on the lift and drag of the kite and varies
within one trajectory period. The kite’s speed is controlled with fixed angle of
attack and variable angle of inclination, which provides the optimum performance,

32
3.2. Tidal Power Plant System Modeling

based on the analysis presented in Sections 5.13 and 5.16 of [54]. Therefore, the
average value of the water speed Va can be calculated as follows [47]

2 L
Vamean = Vtide (3.14)
3 Dk
where L is the kite lift coefficient, Dk is the drag coefficient and Vtide is the free
stream speed of the tide, which is considered unaffected by the operation of the
kite. The Vtide varies based of the lunar cycle, as described in Section 2 of [54]. In
case the kite’s trajectory in the sea is controlled with fixed angle of attack and
variable angle of inclination, the optimal lift to drag ratio is equal to DLk = 6.512,
as stated in Fig. 36 of [54], and the water speed speed in (3.14) becomes

[54] 2
Vamean == · 6.512 · Vtide = 4.34 · Vtide (3.15)
3
The effective water speed Va varies periodically around its calculated mean value,
since the kite is moving at some time instants with or against the tidal current,
as shown in Fig. 3.3(a). These water speed fluctuations create a variation of
the extracted mechanical power, as well, since in (3.6) it can be seen that the
mechanical power is a function of the water speed. Therefore, the turbine rotational
speed ωturb and shaft torque vary also periodically based on (3.11) and (3.13),
when the turbine has constant T SR and CP .
Experimental measurements of the CP and the T SR for the turbine used in the
TUSK plant are not available. Therefore, the curve shown in Fig. 3.4(a) is
used for the analysis of this chapter, which is referenced from [107–109] where

Table 3.1: TUSK Design Parameters

Parameter Value Unit


Cylinder radius r 25 m
Sphere radius R 100 m
Turbine radius RT 0.75 m
Altitude angle D 45 °
Maximum CP max 0.45 -
Optimal T SRopt 6.3 -
Power plant base power Sb 500 kVA
Turbine base speed ωturb b 84 rad/s
Maximum effective water speed Va max 10 m/s

33
Chapter 3. Modelling and Control of Kite-based Tidal Power Generator Systems

10 0.8
Vtide= 2 m/s
Vtide= 2 m/s
9 X: 0.5027
Y: 8.361 Vtide= 1.9 m/s
0.6
Va [m/s]

Pm [pu]
8 X: 12.57 Vtide= 1.9 m/s
Y: 7.942 0.4
7
X: 1.131
Y: 6.689 0.2
6 Vtide= 1.6 m/s
Vtide= 1.6 m/s
5 0
0 /2 1.5 2 2.5 3 3.5 4 0 /2 1.5 2 2.5 3 3.5 4
Position Parameter t [rad] Position Parameter t [rad]
(a) (b)
Fig. 3.3: (a) Effective water speed in m/s and (b) turbine power in pu for maximum CP
as a function of the kite’s position on the Viviani’s curve

X: 6.3
0.5 Y: 0.45

0.4

0.3
CP

0.2

0.1

0
0 2 4 6 8 10 12
TSR
(a)
Maximum Power Pgrid [pu]

0.8 Vtide = 2 m/s

0.6 Vtide = 1.9 m/s

0.4
Vtide = 1.6 m/s

0.2
TSR = const.= 6.3
0
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Optimal Rotor speed turb
[pu]
(b)
Fig. 3.4: (a) CP curve for marine current turbine as referenced from [107–109]; (b)
Grid power in pu as a function of the turbine rotational speed for optimal
T SRopt = 6.3 and maximum CP max = 0.45

marine current turbines are studied. When optimal T SRopt = 6.3 and maximum
CP max = 0.45 are considered, the equations (3.11) and (3.13) are converted into

34
3.2. Tidal Power Plant System Modeling

3
Pgrid maxi = k1 · ωturbi
− Plossi − Pe lossi
2 Plossi Pe lossi (3.16)
Tgrid maxi = k2 · ωturb − −
i
ωturbi ωturbi

The TUSK plant’s design characteristics of Table 3.1 are used as base model for
the power calculations of Fig. 3.3 and Fig. 3.4. A sample plant with 500 kW
nominal turbine power has been chosen for the analysis of this section. Fig. 3.3(a)
shows the fluctuation of Va and Fig. 3.3(b) shows Pm during one period of the
kite’s motion for tidal stream speeds equal to 1.6 m/s, 1.9 m/s and 2 m/s. The
dashed lines in Fig. 3.3(a) show the average effective water speed of the tidal
stream speed. The average extracted power is higher, when the tidal stream
increases; however, the ripple in the extracted power increases, as well. Fig. 3.4(b)
shows a graphic depiction of (3.16), the grid power. For a given tidal stream speed
Vtide the extracted power varies around a middle point, since the kite’s speed is
changing based on its periodic movement.
The operating region for turbine speed ωturb higher than the nominal value of
1pu has not been considered at this study, due to mechanical constraints of the
experimental set-up in Section 3.3. In the real-life kite system the turbine speed
would be limited by the pitch controllers, after it exceeds the nominal value.

3.2.2 MPPT with Active Control of Generator Rotor Speed


An MPPT control strategy where the rotational speed of the generator is directly
controlled is shown in this subsection. The control system of the generator-side
VSC consists of an MPPT block that generates the reference rotor speed ωgen ∗
in
rad/s, an outer speed controller loop and an inner current controller, which finally
generates the reference voltage v∗gen αβ of the inverter. The control structure of the
SG is based on the FOC strategy for synchronous machines, previously described
in Section 2.2. The block diagram of the MPPT control can be seen in Fig. 3.5.
The MPPT algorithm presented in [110], which is widely used in wind power
plants, has been adopted here in order to develop a suitable control scheme for
TUSK power plants.
More specifically, the active power at the grid side of the drive is measured
and the optimal rotor speed is calculated though (3.11) for T SR = T SRopt and
CP = CP max in order to maximize the power production of the plant. A low-pass
filter is applied on the reference rotor speed to filter out high frequency harmonics
that come from the grid power measurement.

35
Chapter 3. Modelling and Control of Kite-based Tidal Power Generator Systems

MPPT
Pgrid(ωturb)
[Fig. 3.4(b)] LPF

Pgrid igen d igen q


measured Generator Speed Controller
grid power TSRopt
KI ω 1/s Saturation
 gen
* v*gen 
+ + * *
igen Generator 3-phase
- + Tgen q
ωgen -1 KP ω 1/ψmd Current SVM Gen-side
gen.
mode + Controller VSC
Βα ω *
igen d =0A

Fig. 3.5: Block diagram of the synchronous generator FOC-based control system with
separate speed control loop

The T SRopt and CP max in (3.11) depend on the geometry of the turbine and are
obtained through experimental results in practical applications. In this study,
the tidal turbine optimal T SRopt and maximum CP max are equal to the values
provided in Table 3.1, which have been referenced from [107–109].

3.2.3 MPPT with Torque Control of the Generator


Another way of operating the turbine on the maximum power point is by generating
directly the necessary electrical reference torque without the use of any speed
controller loop, as referenced from [111]. There is no active control of the shaft
speed, since the speed is only a result of the balancing between the generator
electrical torque and the mechanical torque of the turbine. The main advantage
of this MPPT algorithm is the simplicity of the generator control, since only the
current controller is needed. An analytical comparison of both MPPT methods is
described in Subsection 3.4.2.
Input of the MPPT control algorithm is the measured rotor speed and the optimal
electrical torque is calculated based on the torque-speed equation in (3.13) for

igen d igen q
MPPT Current ref. generation
Tgrid(ωturb)
(eq. 3.13) LPF Saturation v*gen 
*
Tgrid + *
Tgen *
igen Generator 3-phase
q
ωgen 1/ψmd Current SVM Gen-side
+ Controller VSC
TSRopt
Plosses × *
igen d
[Fig. 3.9(c)] =0A
÷
ωgen

Fig. 3.6: Block diagram of the synchronous generator FOC-based control system with
generator torque control

36
3.3. Design of Tidal Power Laboratory Emulator

T SR = T SRopt and CP = CP max , so that the turbine can operate under MPPT.
The torque given by (3.13) is the electrical torque at the grid side of the kite
and, therefore, the losses of the system have to be added in order to calculate the
reference torque of the generator.
Pelectrical loss

Tgen ∗
= Tgrid + (3.17)
ωgen

The losses Pelectrical loss include the stator losses of the generator, the B2B converter
and the grid filter losses and are shown in Fig. 3.9(c). The reference current Iq∗ is
calculated by dividing the reference torque with the flux of the generator and the
reference current is fed to the generator current controller. The block diagram of
the control system for the generator-side converter is shown in Fig. 3.6, when the
generator is controlled in torque control mode.

3.3 Design of Tidal Power Laboratory Emulator


A tidal kite emulator was designed in order to test the system model described
in Section 3.2 and the control strategy of the SG. Fig. 3.7 shows the structure
of the experimental set-up, which consists of an IM and an externally excited
synchronous generator. The SG is driven by B2B-connected VSCs, which operate
similarly to the generator drive of a real tidal power plant.
The generator-side VSC controls the generator voltage and speed using the field-
oriented control scheme [112] (see Section 2.2) and voltage-oriented control is used
for the grid-side VSC [102] (see Section 2.3). In order to simplify the machine
control, it is set i∗gen d = 0 A and the generator torque is controlled only by the
q-axis current. There is a grid filter Lf at the output of the grid-side VSC and a
resistor bank Rinit for soft-charging the dc-link capacitors at the initial start-up

If DC-link
Filter DC-link Filter
Grid (+) Grid
3-phase SW3 3-phase 3-phase (+)
SW1
L1 L2 3-phase Lf Rinit
Cf Diode Motor- IM SG Gen-side
Rdc
Grid-side
side VSC VSC VSC
400V Rect. SW2 400V
igen is vs Pgrid
θgen ωgen Vdc tot
* (time)
TeIM
IM Direct Torque Generator Grid Converter
Control ±1500 Control System Control System
Vtide Va(time) rpm (Fig. 3.5/3.6 and 2.3) (Fig. 2.4 and 2.5)
eq. * *
Pgrid (time) + PeIM (time) ×
TUSK Cp(TSR) (3.11)
RT Tidal (3.9) ωgen
(Fig. 3.4a) + ÷
Turbine Pe loss ωgen Pgrid MPPT
R Pm(ωturb)
Model (Fig. 3.9a)
Cp max , TSRopt (Fig. 3.5/3.6)
Cp max , TSRopt
r TUSK Emulator (Fig. 3.4b) Tidal Generator
(Fig. 3.4b)

Fig. 3.7: Power circuit and control system diagram of the TUSK tidal plant emulator

37
Chapter 3. Modelling and Control of Kite-based Tidal Power Generator Systems

Power converters of
the SG drive

Grid filter Lf

(a)

IM
Flywheel SG
(turbine emulator)

(b)

Fig. 3.8: Laboratory set-up of a 35 kVA converter-driven TUSK tidal plant emulator:
(a) overview of the set-up and (b) the electrical machines test bench with the
flywheel installed

of the system. The dc-link of the generator drive is protected from overvoltage
with a braking chopper Rdc .
The IM is supplied by a VSC and a passive diode rectifier and emulates the behavior
of the TUSK tidal turbine. The control system of the SG power converters is
implemented on the dSPACE 1005 control board. The design parameters of the
experimental emulator and the base of the experimental analysis per-unit system
are presented in Table 3.2. An overview picture of the laboratory set-up is shown
in Fig. 3.8(a).
In order to emulate the turbine shaft inertia of a real TUSK tidal plant, a flywheel

38
3.3. Design of Tidal Power Laboratory Emulator

Table 3.2: Tidal Power Emulator Design Parameters

Parameter Value Unit


Base rotor speed nb 1500 rpm
Base power Sb 35 kVA
Base current Ib 71.7 A
SG number of pole pairs P 2 -
Shaft inertia (without the flywheel) J 2.04 kg·m2
Shaft inertia (with the flywheel) J 0 7.78 kg·m2
SG friction constant b 0.073 Nm/rad·s
SG excitation If 22 A
SG Stator resistance Rs 60 mΩ
SG Stator inductance Ld 10.8 mH
SG Stator inductance Lq 6.2 mH
SG d-axis mutual inductance (for rated If ) Lmd 10.4 mH
Power converters’ switching freq. fsw 5 kHz
Grid filter inductance Lf 2 mH
DC-link capacitance 6.08 mF
Nominal dc-link voltage of the SG drive Vdc tot 670 V
Nominal L2L grid voltage (rms) Vs 400 V

has been installed in the machine set-up as shown in Fig. 3.8(b). The flywheel has
the shape of a circular disc and a mass of approximately 90 kg. The additional
inertia that the flywheel adds on the shaft of the electrical machines is equal to
5.74 kg · m2 .
The IM of the experimental set-up is controlled in Direct Torque Control (DTC)
mode by an ACS600 ABB drive [113]. The rated rotational speed of the IM is
1445 rpm. The electromagnetic torque reference TeIM∗
is calculated by the tidal
turbine model, which has been previously presented in Section 3.2.
As seen in Fig. 3.7, inputs of the tidal turbine model are the design characteristics
of the TUSK trajectory and turbine geometry, while outputs of the model are the
optimal power-rotor speed curve (Fig. 3.4), the maximum CP max , optimal T SRopt
and a time series of the effective water speed Va . Then, the active power of the
TUSK plant is calculated based on (3.11). The losses of the laboratory emulator
are compensated by adding them to the reference grid power Pgrid ∗
(time). Finally,
the reference electromagnetic torque of the IM TeIM is calculated by dividing the

39
Chapter 3. Modelling and Control of Kite-based Tidal Power Generator Systems

active power PeIM



(time) with the rotor speed.
The losses of the laboratory emulator have been measured with the oscilloscope and
are shown in Fig. 3.9 for many different operational points of the SG. Specifically,
the measured losses in Fig. 3.9(a)-(b) include the mechanical losses of the IM and
SG, the core and copper losses of the SG, the generator and grid VSC losses and
the grid filter Lf losses. Fig. 3.9(c)-(d) shows only the converter and grid filter
losses of the SG drive.

5 0.2
Losses [kW]

Losses [pu]
4 18.21 Hz
30 Hz 0.1
3 35 Hz
40 Hz
2 45 Hz 0
50 Hz 1
0.8 1.2
1 0.6 1
0 5 10 15 20 25 30 35 0.4 0.8
0.2 0.6
Grid power [kW] Grid power [pu] Rotor speed [pu]
(a) (b)
2.5
0.1
2
Losses [kW]

Losses [pu]

1.5 30 Hz 0.05
35 Hz
1 40 Hz
45 Hz 0
1
50 Hz 1.2
0.5 0.5 1
0 5 10 15 20 25 30 35 0.8
Grid power [kW] Grid power [pu] Rotor speed [pu]

(c) (d)
Fig. 3.9: Curve-fitted diagram of the TUSK tidal power emulator losses including the
(a)-(b) total SG drive losses and the IM mechanical losses and (c)-(d) the
electrical losses of the power converters and grid filter in the SG drive

3.4 Experimental Results of the Laboratory


Emulator
The experimental results of this section are divided into two parts. In Subsection
3.4.1 an analysis of the bandwidth of the speed controller is made and its effect on
the kite’s power generation is shown. Experimental results from the TUSK plant
emulator are presented, while in this case the flywheel shown in Fig. 3.8(b) is not
installed in the set-up in order to allow faster responses of the speed control.
In Subsection 3.4.2 a comparison of the MPPT algorithms from Subsections 3.2.2

40
3.4. Experimental Results of the Laboratory Emulator

and 3.2.3 is made for different operating conditions of the TUSK. The flywheel is
installed in this case on the shaft of the machine system, in order to study also the
effect of the inertia on the response of the system and to point out any possible
problems that can arise.

3.4.1 Effect of the Speed Controller’s Response Time on


the kite’s Power Generation
The transient response experimental results of the SG speed controller are shown
in Fig. 3.10(a)-(b), where a step-change of the speed reference is performed from
0.6 pu (corresponding to 900 rpm) to the nominal speed of 1 pu (which is equal
to 1500 rpm). As can be seen in the figure, the rise time tr is 0.31 s, when the
speed controller bandwidth is aω = 2π rad/s. The reference q-axis current of
the generator Igenq

reaches also its maximum positive value of 71.7 A, while the
machine is accelerating. Due to the fast acceleration applied in this test, the SG
operates temporarily in motor operation and the current becomes positive. A
slower speed controller with bandwidth equal to aω = 2π · 0.1 rad/s has also been
tested in Fig. 3.10(c)-(d), which has a rise time tr of 3.66 s. The tr is defined as

1 X: 0.1792 I*gen q [A]


X: 0.05 50 Y: 71.94
Shaft Speed [pu]

Y: 1 X: 0.4426 Igen q [A]


0.9 Y: 0.96
Igen [A]

0
0.8
X: 1.926
Y: -39.01
*gen [pu]
0.7
[pu] -50
gen
X: 0.1372
0.6 Y: 0.64
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Time [s] Time [s]
(a) (b)

1 I*gen q [A]
X: 0.05 50
Shaft Speed [pu]

Y: 1 X: 3.905 Igen q [A]


0.9 Y: 0.96
Igen [A]

X: 0.0506
0 Y: -7.988
0.8
X: 9.619
*gen [pu] Y: -38.51
0.7
[pu] -50
gen
X: 0.2486
0.6 Y: 0.64
0 2 4 6 8 10 0 2 4 6 8 10
Time [s] Time [s]
(c) (d)

Fig. 3.10: Experimental results of the SG in a step-change of the reference speed from
0.9pu (900rpm) to 1pu (1500rpm) with speed controller bandwidth equal to
(a)-(b) aω = 2π rad/s and (c)-(d) aω = 2π · 0.1 rad/s

41
Chapter 3. Modelling and Control of Kite-based Tidal Power Generator Systems

the response time of the generator between 10% and 90% of the final value of the
reference speed.
The TUSK tidal plant model shown in Subsection 3.2.1 is tested experimentally on
the laboratory emulator and the power curves of Fig. 3.11 are imported into the
dSPACE control platform. The power and speed calculations have been converted
into the per-unit system in order to scale down the real 500 kW tidal turbine (Table
3.1) into the 35 kVA electric drive of the emulator (Table 3.2). Aim of this test
is to investigate if the generator speed controller can follow effectively the speed
reference given by the MPPT block. The MPPT control algorithm with separate
speed control loop of the generator (Fig. 3.5) is used in this subsection. As can be
seen in Fig. 3.11, a decrease of the Viviani’s curve r radius results to faster turbine
power variations and faster variations of the generator speed reference, since the
system is controlled with T SRopt and CP max . More specifically, r = 25 m results
to a power variation with period 22.9 s, while at r = 5 m the period of the power
is 4.6 s.
The experimental results of the emulator for Viviani’s curve radius equal to
r1 = 25 m and r2 = 5 m are shown in Fig. 3.12. The generator actual speed ωgen
follows the reference shaft speed given by the MPPT control block with small
error, when the bandwidth of the speed controller is equal to aw = 2π rad/s and
the radius of the trajectory is r1 = 25 m [Fig. 3.12(a)]. However, when the tidal
kite’s motion becomes faster at r2 = 5 m, the error between the reference speed
and the actual shaft speed becomes larger. This error increases even more when
the bandwidth of the speed controller is equal to aw = 0.2π rad/s.
The error between the reference rotor speed and the actual speed causes the

0.6 1
TUSK power P*grid [pu]

r = 25 m r = 25 m
0.5 r=5m 0.9 r=5m
*gen [pu]

0.4
0.8
0.3
0.7
0.2

0.1 0.6
5 10 15 20 25 0 5 10 15 20 25
Time [s] Time [s]
(a) (b)
Fig. 3.11: (a) Reference grid power and (b) shaft speed of the tidal power emulator for
modelling a TUSK tidal power plant with tidal stream speed Vtide = 1.9 m/s
and Viviani’s curve radius r1 = 25 m and r2 = 5 m.

42
3.4. Experimental Results of the Laboratory Emulator

Shaft Speed [pu]


0.9

0.8

0.7 *gen [pu]

gen
[pu]
0.6
25 30 35 40 45 50
Time [s]
(a) aω = 2π rad/s, r = 25 m
1
Shaft Speed [pu]

0.9

0.8

0.7 * [pu]
gen
[pu]
gen
0.6
25 30 35 40 45 50
Time [s]
(b) aω = 2π rad/s, r = 5 m
1
Shaft Speed [pu]

0.9

0.8

0.7 *gen [pu]

gen
[pu]
0.6
25 30 35 40 45 50
Time [s]
(c) aω = 0.2π rad/s, r = 25 m
1
Shaft Speed [pu]

0.9

0.8

0.7 *gen [pu]

gen
[pu]
0.6
25 30 35 40 45 50
Time [s]
(d) aω = 0.2π rad/s, r = 5 m
Fig. 3.12: Experimental results of the tidal power emulator reference and actual shaft
speed with different profiles of the TUSK trajectory and different generator
speed controller bandwidths

43
Chapter 3. Modelling and Control of Kite-based Tidal Power Generator Systems

Emulator power [pu]


0.6

0.4
X: 26.12 P*gridOpt [pu]
Y: 0.3448
0.2 Pgrid [pu]
Pgrid mean [pu]
0
25 30 35 40 45 50
Time [s]
(a) aω = 2π rad/s, r = 25 m
Emulator power [pu]

0.6

0.4

X: 34.16 P*gridOpt [pu]


0.2 Y: 0.3126 Pgrid [pu]
Pgrid mean [pu]
0
25 30 35 40 45 50
Time [s]
(b) aω = 2π rad/s, r = 5 m
Emulator power [pu]

0.6

0.4
X: 26.7 P*gridOpt [pu]
Y: 0.344
0.2 Pgrid [pu]
Pgrid mean [pu]
0
25 30 35 40 45 50
Time [s]
(c) aω = 0.2π rad/s, r = 25 m
Emulator power [pu]

0.6

0.4

X: 33.54 P*gridOpt [pu]


0.2 Y: 0.3094 Pgrid [pu]
Pgrid mean [pu]
0
25 30 35 40 45 50
Time [s]
(d) aω = 0.2π rad/s, r = 5 m
Fig. 3.13: Experimental results of the tidal power emulator reference and actual grid
power with different profiles of the TUSK trajectory and different generator
speed controller bandwidths.

44
3.4. Experimental Results of the Laboratory Emulator

emulated power plant to operate with T SR different from the optimal value, which
also results to a decrease in the produced power. Fig. 3.13 shows with dashed
blue line the active power at the grid side that the emulator would produce, if
it could always operate with optimal T SR and maximum Cp . The actual power
Pgrid produced by the emulated plant is also shown with red line and the average
actual power is shown with black dashed line. It can be seen that the produced
power of Fig. 3.13(a) is the closest to the optimal case, while in all the other cases
the actual power is a little lower than the optimal power. There is also a time
shift between the PgridOpt and Pgrid , which is more visible in Fig. 3.13(b) and
Fig. 3.13(d). This delay is introduced by the PI controller of the generator speed
control loop and by the LPF at the output of the MPPT block (Fig. 3.5).

3.4.2 MPPT Algorithm Comparison Study


The MPPT algorithms described in Subsections 3.2.2 and 3.2.3 are now compared
experimentally on the tidal power emulator. The flywheel is installed in this case
on the shaft of the machine system and the inertia of the mechanical system
increases compared to the previous study and is equal to J 0 = 7.78 kg·m2 . The
transient response of the SG speed controller is shown in Fig. 3.14, where a
step-change of the speed reference is performed from 0.6 pu (corresponding to
900 rpm) to the nominal speed of 1 pu (which is equal to 1500 rpm). Bandwidths
of aω = 2π rad/s and of aω = 2π · 0.1 rad/s are tested, similarly to Subsection
3.4.1.
As can be seen in Fig. 3.14(a), the rise time tr is 1.07 s, when the speed controller
bandwidth is aω = 2π rad/s, which is as expected more than three times higher
than in the case without the flywheel. The rise time becomes equal to tr = 3.46 s,
when the slower speed controller aω = 2π · 0.1 rad/s is tested [Fig. 3.14(c)].
The TUSK tidal plant model is tested experimentally on the laboratory emulator
and the power curves of Fig. 3.11 are imported into the dSPACE control platform.
The MPPT control algorithm with speed control loop (Fig. 3.5) is compared with
the algorithm with torque control of the generator (Fig. 3.6) in the following
figures in order to identify which control structure has the best performance.
The experimental results of the shaft speed for Viviani’s curve radius equal to
r1 = 25 m and r2 = 5 m are shown in Fig. 3.15. The blue curve shows the
reference speed ωgen

which is generated by the MPPT block, when the generator
is controlled with speed controller and bandwidth equal to 0.2π rad/s. The black
curve shows the optimal reference speed ωgenOpt

, which is the reference speed that

45
Chapter 3. Modelling and Control of Kite-based Tidal Power Generator Systems

1
X: 0.3708 I*gen q [A]
X: 0.05 50 Y: 72.59
Shaft Speed [pu]

Y: 1 X: 1.32 Igen q [A]


0.9 Y: 0.96

Igen [A]
0
0.8 X: 4.951
Y: -37.68
*gen [pu]
0.7 X: 0.2458
Y: 0.64 [pu] -50
gen
0.6
0 1 2 3 4 5 0 1 2 3 4 5
Time [s] Time [s]
(a) (b)
X: 0.0636
Y: 60.28
1 I*gen q [A]
X: 0.05 50
Shaft Speed [pu]

Y: 1 X: 3.728 Igen q [A]


0.9 Y: 0.96

Igen [A]
0
0.8
X: 9.856
Y: -38.19
*gen [pu]
0.7 X: 0.2652
Y: 0.64 [pu] -50
gen
0.6
0 2 4 6 8 10 0 2 4 6 8 10
Time [s] Time [s]
(c) (d)

Fig. 3.14: Experimental results of the SG with the flywheel installed on the shaft in a
step-change of the reference speed from 0.9 pu (900 rpm) to 1 pu (1500 rpm)
with speed controller bandwidth equal to (a)-(b) aω = 2π rad/s and (c)-(d)
aω = 2π · 0.1 rad/s

the emulator would have if it operated with constant CP max = 0.45. Fig. 3.15(a)-
(b) show the experimental results of the generator when the speed controller is
used, while in Fig. 3.15(c)-(d) the torque control mode is used.
As can be seen in Fig. 3.15, there is a delay ∆t between the actual shaft speed
ωgen and the ωgenOpt

. The delay is higher in the case of the torque control mode
compared to the speed control mode by 0.7 s, when the Viviani’s curve radius is
equal to r1 = 25 m [Figs. 3.15(a) and 3.15(c)] and higher by 0.07 s when r2 = 5 m
[Figs. 3.15(b) and 3.15(d)]. Due to this delay ∆t the active power produced by
the emulated tidal turbine is expected to be lower in the case of the torque control
mode, as will be shown later. Fig. 3.16 shows the CP curves of the emulator for
all the four operating conditions. In all the graphs the CP curve reaches the peak
point of 0.45 when the shaft speed ωgen overlaps with the ωgenOpt

. However, the
slower response of the generator operating in torque control mode is also visible
here.
Fig. 3.17 shows the generated active power at the grid side of the emulator. The
red solid line shows the actual power of the grid Pgrid , the dashed blue line is
the reference grid power Pgrid

and the dashed black line represents the optimal

46
3.4. Experimental Results of the Laboratory Emulator

1X: 27.44
X: 29.74
Y: 0.9294 * genOpt [pu]
Y: 0.9119
* gen [pu]

Shaft Speed [pu]


0.9
gen
[pu]
0.8
t
0.7

0.6
25 30 35 40 45 50
Time [s]
(a) aω = 0.2π rad/s, r = 25 m
1
* genOpt [pu]
X: 28.64
Y: 0.8732 X: 29.7 * gen [pu]
Shaft Speed [pu]

0.9
Y: 0.837
gen
[pu]
0.8

0.7 t

0.6
25 30 35 40 45 50
Time [s]
(b) aω = 0.2π rad/s, r = 5 m
1 X: 27.44
Y: 0.9293 * genOpt [pu]
Shaft Speed [pu]

0.9 gen
[pu]
X: 30.44
Y: 0.8884
0.8
t
0.7

0.6
25 30 35 40 45 50
Time [s]

(c) Torque control, r = 25 m


1
X: 28.64
Y: 0.873
Shaft Speed [pu]

0.9
X: 29.77
Y: 0.8018
0.8

0.7 * genOpt [pu]


t
gen
[pu]
0.6
25 30 35 40 45 50
Time [s]
(d) Torque control, r = 5 m
Fig. 3.15: Experimental results of the tidal power emulator reference and actual shaft
speed with different profiles of the TUSK trajectory and different MPPT
methods

47
Chapter 3. Modelling and Control of Kite-based Tidal Power Generator Systems

0.46 X: 29.53 X: 30.4 0.46 X: 29.46 X: 29.65


Y: 0.45 Y: 0.45 Y: 0.45 Y: 0.45

0.44 0.44

Cp
Cp

0.42 0.42
Torque control Torque control
Speed control (a = 0.2 rad/s) Speed control (a = 0.2 rad/s)
0.4 0.4
25 30 35 40 45 50 25 30 35 40 45 50
Time [s] Time [s]
(a) r = 25 m (b) r = 5 m
Fig. 3.16: Experimental results of the tidal power emulator power coefficient CP with
different profiles of the TUSK trajectory and different MPPT methods.

reference power PgridOpt



that the emulator would produce with constant T SRopt
and CP max . The average power Pgridmean is shown with dashed red line and is
higher when the MPPT with speed control is used in Figs. 3.17(a)-(b) compared
to the results in Figs. 3.17(c)-(d).
Fig. 3.18 shows a closer look of the reference grid powers, where the peak values
of the reference active power can be seen. Higher reference power can be seen here
as well for the MPPT with speed control.
It can be concluded that the generator has a faster dynamic response when the
MPPT algorithm with speed control loop is used compared to the MPPT scheme
with torque control. The result of this faster response is that the generator can
produce higher power at the peak points as well as in average in one cycle of the
kite’s trajectory. The advantageous performance of the generator speed controller
can be explained by considering the time constant of the mechanical model of
the generator. The time constant of the generator, when there is no active speed
control loop and when torque control mode is used, is equal to

J0
τT ctrl = (3.18)
b
The time constant of the generator, when the speed controller is used, which is
described in Subsection 3.2.2, is equal to

KP ω J0
τSctrl = = (3.19)
KIω b + Baω
It is clear that τT ctrl > τSctrl , which explains the previously presented experimental
results.

48
3.4. Experimental Results of the Laboratory Emulator

0.6

Emulator output power [pu]


0.5 X: 29.89
Y: 0.5247
X: 28.46
0.4 Y: 0.3652

0.3 Pgrid [pu]


P* [pu]
grid
0.2 P* [pu]
gridOpt
P [pu]
grid mean
0.1
25 30 35 40 45 50
Time [s]

(a) aω = 0.2π rad/s, r = 25 m


0.5

0.45
Emulator output power [pu]

X: 43.85
0.4 Y: 0.3872

0.35
X: 33.61
0.3 Y: 0.3393
0.25 P [pu]
grid
P* [pu]
0.2 grid
P*gridOpt [pu]
0.15
Pgrid mean [pu]
0.1
25 30 35 40 45 50
Time [s]

(b) aω = 0.2π rad/s, r = 5 m


0.6
X: 30.1
Y: 0.5212
Emulator output power [pu]

0.5

0.4 X: 29.03
Y: 0.3507

0.3 Pgrid [pu]


P*grid [pu]
0.2 P*gridOpt [pu]
Pgrid mean [pu]
0.1
25 30 35 40 45 50
Time [s]

(c) Torque control, r = 25 m


0.5

0.45
Emulator output power [pu]

X: 43.9
0.4 Y: 0.3724
0.35
X: 33.83
0.3
Y: 0.336
0.25 Pgrid [pu]
P* [pu]
0.2 grid
P*gridOpt [pu]
0.15
Pgrid mean [pu]
0.1
25 30 35 40 45 50
Time [s]

(d) Torque control, r = 5 m


Fig. 3.17: Experimental results of the tidal power emulator reference and actual grid
power with different profiles of the TUSK trajectory and different MPPT
methods

49
Chapter 3. Modelling and Control of Kite-based Tidal Power Generator Systems

X: 27.92
X: 27.92
0.6 Y: 0.5868
Y: 0.5787

Emulator output power [pu]


X: 27.92
0.55
Y: 0.5723

0.5

0.45
P*grid [pu] Torque control
0.4 P*grid [pu] Speed control at a = 0.2 rad/s
P*gridOpt [pu]
0.35
20 22 24 26 28 30 32 34
Time [s]

(a) r = 25 m
0.5 X: 28.71
Y: 0.4841 X: 28.73
Emulator output power [pu]

0.48 Y: 0.4722

0.46 X: 28.73
Y: 0.4636
0.44

0.42
P* grid [pu] Torque control
0.4 P* grid [pu] Speed control at a = 0.2 rad/s
P* gridOpt [pu]
0.38
26 27 28 29 30 31 32
Time [s]

(b) r = 5 m
Fig. 3.18: Experimental results of the tidal power emulator reference grid power

3.5 Summary
Chapter 3 presents the model of a TUSK tidal power generator unit describing
the kite’s trajectory and the produced electric power. It has been shown that
the tidal turbine produces fluctuating electrical power due to the motion of the
TUSK into the sea. Two different algorithms of MPPT have also been presented,
one with closed loop speed control and the second method with torque control
of the generator. A laboratory emulator of the TUSK has been constructed for
reproducing the performance of the tidal system in laboratory conditions. The
accuracy of the shaft speed control has been evaluated with measurements on the
emulator for different trajectories of the TUSK and gains of the speed controller.
The experimental results are divided into two parts. The first part in Subsection
3.4.1 analyzes the effect of the speed controller bandwidth on the performance of
the MPPT. Specifically, Fig. 3.13 shows that an increase of 10 times of the PI
controller gains results to an increase of the average produced power of up to 1%,
since the speed controller of the generator converter can follow the reference rotor
speed with smaller error. Therefore, a better design of the PI controllers regulating
the converters and a fast speed controller can increase the power produced by the
TUSK tidal power system.

50
3.5. Summary

The second part of the experimental results in Subsection 3.4.2 compares two
MPPT control schemes, the MPPT with torque control of the generator which
is simpler and easier to design and the control scheme with active speed control.
The speed control method has advantageous dynamic response which results to
a higher power production of the tidal turbine, although the difference in the
performance of both methods is quite small in some cases. Specifically, the average
power production of the generator Pgrid mean is approximately 4% higher when
using active speed control compared to the MPPT with torque control, as shown
in the Figs. 3.17(a) and 3.17(c) where it increases from 0.3507 pu to 0.3652 pu.
The results of this chapter can be utilized to improve the power production
of tethered kites. The advantages and disadvantages of each MPPT control
scheme need to be considered in order to evaluate if the more efficient, but also
more complicated, speed control method should be used or the torque control
method is better due to its simplicity. The experimental set-up of the tidal
power emulator can also be used for modelling the power generation from other
renewable energy sources, such as the wind power. The wind speed variations are
stochastic, contrarily to the kite tidal speed variations which have been studied
here. Therefore, emulating wind speed variations with the IM of the current
experimental set-up could be of high value in order to investigate and optimize
the performance of the generator.

51
Chapter 3. Modelling and Control of Kite-based Tidal Power Generator Systems

52
Chapter 4

Multilevel Converter Topologies and


Modulation Techniques

4.1 Motivation towards Medium Voltage Drives


Increasing the operating voltage of the power generation system has attracted
considerable research interest in offshore applications, since it can lead to higher
power density. Electrical machine drives with higher voltage have lower current
flow in the cables and the power converters, which reduces copper loss and leads
to a possible reduction of the cables size. Medium voltage drives are particularly
popular in power generation from renewable energy sources (RES), including wind
and tidal/wave power systems, where the high power density can lead to a decrease
of the system level cost. Large power generation units (e.g. 10 MW or even higher)
suffer from high current on the cables and on the semiconductors of the converters
[91], whereas by increasing the operating voltage of the system, the conduction
losses can be considerably reduced.
Especially in the case of the TUSK, the power cables, which are installed inside
the tether of the kite, can hinder its movement inside the sea and create additional
drag which is reflected in the power production equation (3.6). Therefore, it
is of high importance to decrease the size and weight of the cables as much as
possible. Also, the kite is connected to the onshore PCC through long cables
which lay on the sea-bed. The power losses and the voltage drop through these
cables need to be kept at minimum level in order to connect to the local grid.
Increasing the operating voltage of the power generation system will decrease the

53
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

copper losses through these undersea cables. Therefore, the efficiency of the power
transportation system can be highly increased, if the copper area of the cables is
kept constant. Otherwise, the cable size can be reduced making the kite system
more compact, if it is decided the efficiency of the cables to remain the same as
with a low-voltage system.

4.2 Multilevel Converter Topologies & Capacitor


Voltage Balancing
The above issues can be effectively addressed by using medium voltage multilevel
converters in the generator drive. The basic concept of these topologies is the
interconnection of many low-voltage-rated power switches. A simple series connec-
tion of power switches may increase the operating voltage of the converter as that
shown in the traditional 2-level converter (2LC) of Fig. 4.1(a). However, this type
of converter faces the serious problem of unbalanced voltage sharing among the
power switches due to the system parasitic components and the uneven charac-
teristics of the power switches. This problem can be solved by using balancing
circuits or special designed gate drivers that ensure synchronous turn on/off of
the switches [66, 67]. The losses of the medium voltage 2LC are also higher than
in the multilevel topologies described later [67].
Multilevel converters can be defined as the inverters that generate at least three
voltage levels at their ac side. Multiple low-voltage-rated power switches are used
in these topologies, so that the converter is able to operate with high voltage. This
topology also attains reduced total harmonic distortion at the ac side, compared
to the 2-level VSC. This is a further advantage in the case of the undersea kites
because the passive filters placed inside the kite at the grid-side of the generator
drive can be drastically minimized saving space and weight. Another advantage
of the multilevel converters is the reduced dV dt of their generated ac voltage
1
,
which is not further studied in this thesis.
Modular converters, such as the Cascaded H-Bridge (CHB) converter or the
Modular-Multilevel-Converter (MMC) [114], are commonly used topologies in
high-voltage applications. These converters due to their modular structure can
have large number of levels handling many tens to hundreds kVs of voltage and,
1 C. Larsson, A. Rydgård, G. Mademlis, Y. Liu, and M. Fredriksson, "Overvoltage Mitigation

of Medium Voltage Electric Drives with Long Cables using Multilevel-Converters and Passive
Filters", in Proc. of the 21st European Conference on Power Electronics and Applications
(EPE’19 ECCE Europe), 2019, P.1-P.10.

54
4.2. Multilevel Converter Topologies & Capacitor Voltage Balancing

therefore, they are commonly used in grid applications for improving power flow
control and grid voltage stability, such as in parallel connection to the transmission
lines as STATCOMs [115] and in HVDC applications [116]. One disadvantage
of these topologies is the lack of a common dc-link at their dc-side, since each
converter module has its own capacitor. Therefore, they are used less frequently
in electric drives where two of these converters need to be connected back-to-back
(B2B), because disturbances at the machine-side can affect more easily other
control variables at the grid-side of the system and vice versa [117]. Also, the need
for many passive components, such as capacitors at each converter module and
inductors for each converter leg, can make the converter bulky.
The Neutral-Point-Clamped (NPC) (also known as Diode-Clamped) converter,
originally introduced in [118], and the Flying-Capacitor (FC) converter are two of
the main multilevel VSC types with a controllable dc-link at their dc-side. Their
3-level [Fig. 4.1(b)] variant has already been used in industrial applications, such
as in [119, 120] for the NPC converter. Both the FC and the NPC converters are
limited up to 5-levels in practical applications, because their structure becomes
too complicated for higher number of levels. The power switches used in these
converters are rated to 1/4 of the total dc-link voltage. Therefore, by using power
switches of 1.2 kV rated voltage, dc-link voltage of up to 4 kV can be obtained
in their 5-level version, depending on the safety voltage margin that we choose
to have for the power semiconductors. The schematics of 5-level FC and NPC
converters are shown in Fig. 4.1(c)-(d).
The main advantage of the FC converter compared to the NPC is the reduced
number of semiconductor components. They have equal number of power switches,
but the NPC converter has also got clamping diodes. This makes the structure
of the FC much simpler compared to the NPC. However, a large number of
capacitors are needed for the FC converter, which becomes even larger in the
5-level version and increases a lot the size of the power conversion system. Proper
balancing of these flying capacitor voltages and precharging them during the system
initialization is needed making the operation of the converter more complicated
[121].
On the other hand, the NPC converter requires the highest number of semiconduc-
tors due to the extra clamping diodes that it has, which also do not have equal
reverse voltage rating. For this reason, series connection of two diodes are needed
in some points, as can be seen in Fig. 4.1(d). However, the NPC converter has
the lowest number of capacitors [122], which keeps the volume of the converter
small, if a high-power-density design is implemented with low loss semiconductor

55
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

devices. Since the size of the power conversion system is of high importance for
the TUSK application, the 5-level NPC inverter is chosen as the preferred VSC
topology studied in the rest of this chapter.
Another serious issue of the NPC converter is the control of its dc-link capacitors.
In general, an N -level NPC converter uses a single dc-bus which is subdivided
into (N − 1) series connected capacitors, whose voltages should be monitored and
properly controlled in order to be balanced in every operating condition.
More specifically, the 3-level NPC converter has two dc-link capacitors which have
normally balanced voltages, when a symmetrical three-phase load is connected at

Phase c Phase c
Phase b Phase b
Phase a Phase a
T1
C2
C2 T1 VC2
T2
c
D1
c
0 b 0
b
Vdc tot
a Vdc tot
T3
a
C1 D2
C1 T2 VC1
T4

(a) (b)

Phase c
Phase b
Phase a
T1
Phase c Phase c C1
Phase b Phase b
Phase a Phase a T2
T1 T1
T3 Submodule
C4 C4 VC4
T2 T2 C2
D1
T4
T3 T3
D3a c
C3 C3 VC3 L1
T4 c D3b T4 c Vdc tot b
Cc3 Cc2 Cc1 b D5a b a
Vdc tot 0 a
0 D5b a T5
T5 Vdc tot D2a T5 C3 L2
C2 C2 VC2 D2b
T6 D4a T6 T6
D4b
T7 T7 T7
C1 C1 V
C1 D6 C4
T8 T8
T8

(c) (d) (e)


Fig. 4.1: Schematic of (a) 2-level VSC with multiple series connected switches, (b) 3-level
NPC, (c) 5-level FC, (d) 5-level NPC and (e) 5-level MMC converters

56
4.2. Multilevel Converter Topologies & Capacitor Voltage Balancing

the ac side. They tend to unbalance only at asymmetrical loads and faults. Many
control techniques have been proposed in the technical literature for alleviating
successfully the above drawback of the 3-level NPC converter and the majority
uses zero-sequence voltage for balancing these voltages [80, 102]. The 5-level NPC
converter, shown in Fig. 4.1(d), may be used with higher operating voltage and the
harmonic distortion at the ac output is lower compared to that of the 3-level NPC.
However, the dc capacitor voltage drift is more apparent here, due to the voltage
unbalance in the four capacitors that may be observed even with symmetrical
three-phase loads. Therefore, a more advanced control method of the dc-link
capacitor voltages needs to be implemented as will be shown in the following
sections.
Unequal loss distribution among the switches of the NPC converter is also a serious
drawback of the NPC converter. This problem can result to asymmetrical loss
distribution and junction temperatures of the semiconductors [121, 123]. The
semiconductor losses can be balanced by replacing the clamping diodes with active
switches and this modified NPC converter is called active NPC (ANPC) [124].
However, the loss distribution issue has not been covered in this thesis.
This chapter introduces the SVM technique properly adjusted for N -level multilevel
converters. Afterwards, simulation results are presented from grid-connected

Table 4.1: Parameters of the simulated 3- and 5-level generator drive

Parameter Value Unit


Nominal Power Pnom 50 kW
Total dc-link voltage 4 kV
Maximum ac L2L voltage Vgen (rms) 2.8 kV
24 (3-level drive)
Number of MOSFETs (each NPC converter)
24 (5-level drive)
12 (3-level drive)
Number of clamping diodes (each NPC converter)
30 (5-level drive)
VDSmax (SCH2080KE MOSFET) 1.2 kV
IDmax (for the MOSFET at 25°C) 40 A
RDSon (MOSFET) 80 mΩ
Peak Reverse Voltage VRRM (C3D10170H diode) 1.7 kV
Continuous forward current (for the diode at 25°C) 29 A
Total dc-link capacitance 100 µF
Switching frequency fsw 20 kHz

57
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

generator drives with 3-level converters and with 5-level NPC converters for the
same operating voltage. The generator of the system is a PMSG, which is controlled
with field-oriented control, while a back-to-back connected grid converter uses the
voltage-oriented control scheme, as described in Chapter 2. The available methods
for balancing the dc-link capacitor voltage in the 5-level drives are also presented.
A comparison of the performance of the different designs of the generator drives
is made in terms of simulated efficiency and voltage-current quality, as well as
number of components required for manufacturing the power converters in each
case.
The design characteristics of the two converters and a list of the minimum number
of components, so that both converter types can operate with the same dc-link
voltage, are shown in Table 4.1.

4.3 Multilevel PWM Modulation Strategies


The PWM modulator of a three-phase dc/ac converter is responsible for trans-
lating the reference voltage signal, created by the current controller, into the
corresponding switching function for each active switch of the converter. The
most commonly used modulation strategies for the 2LC are the sinusoidal PWM
(SPWM) and the Space Vector modulation (SVM). These two methods can be
extended for multilevel converters, as described in the following subsections for
5-level and N -level NPC converters.

4.3.1 Sinusoidal PWM


The SPWM technique is built with a sinusoidal reference signal for each phase
of the converter running with fundamental frequency which is being compared
to a high-frequency triangular carrier. For the 2LC, when the reference signal is
higher than the triangular carrier, the upper switch of the corresponding phase
leg is conducting, whereas the lower switch is off. Oppositely, when the reference
signal is lower than the triangular carrier, the lower switch is conducting and the
upper switch is turned off.
This concept can be extended for an N -level converter by using (N − 1) carriers
shifted in amplitude instead of a single triangular carrier. This means that for
a 5-level converter 4 carriers are needed, as shown in Fig. 4.2. There are many
methods of placing the different carriers depending on their phase sequence, as
presented in [125, 126]. The case shown in Fig. 4.2, where all the carriers are in

58
4.3. Multilevel PWM Modulation Strategies

phase is called PH-disposition [125].


The reference phase voltage of the converter is Vph ∗
(t) = m(t) Vdc2 tot and m(t) is
the three-phase modulation signal. Its peak value is called modulation index m̂
and can get values within m̂ ∈ [0, 1]. The three-phase modulation signal equations
are defined as

ma (t) = m̂ · cos (ωt) (4.1)



mb (t) = m̂ · cos (ωt − ) (4.2)
3

mc (t) = m̂ · cos (ωt − ) (4.3)
3
where ω is the fundamental frequency of the converter. The maximum amplitude of
the reference phase voltage of the converter for this specific modulation technique
is equal to
Vdc (N − 1)

V̂phmax = (4.4)
2
where Vdc is the mean dc voltage of the capacitor cells, which is defined as

Vdc tot = Vdc (N − 1) (4.5)

and Vdc tot is the total dc-link voltage of the N -level NPC converter.
It is possible to extend the range of the voltage and utilize better the dc-link
by injecting third-harmonic components in the voltage reference. Then, the
modulation signal for phase A is defined as [102]
3 2 3
m0a (t) =
m̂ · cos (ωt) − m̂ · cos (ωt) (4.6)
2 3
The term at the right side of the above equation is the third-harmonic component
of the voltage, which is cancelled out in the L2L voltage. However, the maximum

1 1

0.5 0.5
mabc
mabc

0 0
ma ma
-0.5 -0.5 mb
mb
mc -1 mc
-1
0 0.01 0.02 0.03 0.04 0 0.01 0.02 0.03 0.04
Time [s] Time [s]
(a) (b)
Fig. 4.2: Modulating signals corresponding to (a) conventional 5-level SPWM and (b)
third-harmonic injected 5-level SPWM

59
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

amplitude of the voltage is now extended by almost 15% and the modulation
index in (4.6) can get values within m̂ ∈ [0, √23 ] without reaching overmodulation.
Therefore, the maximum amplitude of the reference phase voltage of the converter
becomes
Vdc (N − 1)

V̂phmax = √ (4.7)
3
The switching sequence for the switches of the 5-level NPC converter shown in
Fig. 4.1(d) can be generated based on the algorithm in Table 4.2.
The multicarrier SPWM has the main advantage of being simple and easy to use.
However, the main issue with this modulation strategy is the difficulty to obtain
the multiple carriers with this specific amplitude shift using a regular MCU. It
might be necessary to have an external customized control board which creates
the required carriers and generates the suitable switching signal for each power
switch of the converter. Also, the Space Vector Modulation (SVM) shown in the
next subsection allows more freedom in choosing the switching states, than the
SPWM, which can be utilized for control purposes.

Table 4.2: Switching States for phase A of 5-level NPC Converter with SPWM

Condition State Pulses∗


ma > Vtr1 ∧ ma > Vtr2 ∧ ma > Vtr3 ∧ ma > Vtr4 4 [1 111000 0]
ma < Vtr1 ∧ ma > Vtr2 ∧ ma > Vtr3 ∧ ma > Vtr4 3 [0 111100 0]
ma < Vtr1 ∧ ma < Vtr2 ∧ ma > Vtr3 ∧ ma > Vtr4 2 [0 011110 0]
ma < Vtr1 ∧ ma < Vtr2 ∧ ma < Vtr3 ∧ ma > Vtr4 1 [0 001111 0]
ma < Vtr1 ∧ ma < Vtr2 ∧ ma < Vtr3 ∧ ma < Vtr4 0 [0 000111 1]

for the power switches [T1 T2 T3 T4 T5 T6 T7 T8 ], where 1 = closed, 0 = open

4.3.2 Space Vector Modulation


The SVM allows the direct calculation of the switching state for all the switches
of the converter having as input the amplitude and the angle of the ac voltage
reference. The SVM diagram for a 5-level converter can be derived from the
conventional 2-level space vector plane, as can be seen in Figs. 4.3(a)-(b). The
N -level SVM has 6(N − 1)2 triangles and [N 3 − (N − 1)3 ] voltage vectors [127].
In many points of the space vector plane there are multiple switching states listed
the one underneath each other, which are equivalent and create the same L2L
voltage. These states are called redundant states and can be utilized for control
purposes, such as for balancing the dc-link voltage for the NPC converter.

60
4.3. Multilevel PWM Modulation Strategies

The switching states can be translated into switching pulses and the conversion
is unique for each converter topology. The conversion table for the 5-level NPC
converter has been shown in Table 4.2.
The modulation index of the space vector modulation m̂ is defined as follows

√ ∗

V̂ph 3 N =5 V̂ph 3
m̂ = =⇒ m̂ = (4.8)
(N − 1)Vdc 4Vdc

It can be seen from the above equation that the maximum voltage obtained by
the SVM without overmodulation (for m̂ = 1) is equal to the voltage calculated
by (4.7), where SPWM with third-harmonic injection is used. Fig. 4.3(c) shows
the vector plane of the SPWM without third-harmonic injection (with green color)
and SVM (with red color), as well as the overmodulation region for m̂ > 1 (marked
with yellow color) [128]. However, the overmodulation region of the SVM has not
been studied in this thesis.
The space vector plane is divided into six sectors depending on the angle θ ∈ [0, 2π]
of the reference voltage V̂ph

, as seen in Figs. 4.3(b)-(c). The sector number is
determined by
Sector = ceil(θ · 3/π) (4.9)

The angle θ is rotated based on the sector number as described in Table 4.3, so
that the voltage vector is converted into an equivalent vector of Sector 1.
The normalized value of the voltage V̂ph

can be calculated by


3V̂ph
v∗ = (4.10)
2Vdc

and then it can be transformed into a 60°-coordinate system (α0 β 0 ), as can be seen

Table 4.3: Sector Selection of SVM

Sector Angle θ New θ0 Basic Vector ua Basic Vector ub


[Phasea Phaseb Phasec ] [Phasea Phaseb Phasec ]
1 (0, π/3] θ0
=θ [1 0 0] [1 1 0]
2 (π/3, 2π/3] θ0 = −θ + 2π/3 [1 1 0] [1 1 0]
3 (2π/3, π] θ0 = θ − 2π/3 [0 1 0] [0 1 1]
4 (π, 4π/3] θ0 = −θ − 2π/3 [0 0 1] [0 1 1]
5 (4π/3, 5π/3] θ0 = θ + 2π/3 [0 0 1] [1 0 1]
6 (5π/3, 2π] θ0 = −θ [1 0 0] [1 0 1]

61
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

2nd Sector
040 140 240 340 440

rd
3 141 241 341 441
1st
041 430
Sector 030 130 230 330 Sector
242 342 442
142 431
042 131 231 331 420
031 320
Vˆph*
020 120 220
343 443
243 432 410
143 232 332 421
043 132 321
032 121 221 310
021
344 010
444 110 433
210 θ
244 333 422
144 233 322 411
133 222 311 400
033 122 211 300
022 111 200
011 334 434 100
234 000 423
134 223 323 412
034 123 312 401
023 112 212 301
012 201
001 101
224 324 424
124 413
024 113 213 313 402
010 110 013 302
002 102 202

4th 014
114 214 314 414
403
6th
111 003 103 203 303
011 100 Sector Sector
000

004 104 204 304 404


001 101
th
5 Sector
(a) (b)
β
Overmodulation
in SVM

V  N  1
SPWM Vˆph max  dc
SVM 2

Vˆph V  N  1
Vˆph max  dc
3
θ
α

2  Vdc  N  1
3

Vdc tot  ( N  1)Vdc

(c)

Fig. 4.3: Space vector plane for (a) 2LC and for (b) 5-level VSC; (c) voltage vector plane
indicating the maximum voltage reference for SPWM without third-harmonic
injection (marked in green) and SVM (marked in red)

62
4.3. Multilevel PWM Modulation Strategies

in Fig. 4.4 [127]



vα∗ 0 = v ∗ cos θ0 − v ∗ sin θ0 / 3
2 (4.11)
vβ∗ 0 = v ∗ sin θ0 · √
3

The space vector vα∗ 0 vβ∗ 0 is now decomposed into the following vectors
 

floor(vα∗ 0 ) floor(vβ∗ 0 )
   
VD
 VE   ceil(vα∗ 0 ) floor(vβ∗ 0 ) 
= (4.12)
   
floor(vα∗ 0 ) ceil(vβ∗ 0 )
 
 VF   
VG ceil(vα∗ 0 ) ceil(vβ∗ 0 )

The vectors of (4.12) are multiplied with the basic vectors [ua ub ] of Table 4.3 in
order to convert them back into their original sector
 0   
VD VD " #
 V0   V  ua
 E   E 
 0 = · (4.13)
 VF   VF  ub
VG0 VG

Each edge of the triangles on the space vector plane seen in Fig. 4.3 corresponds
to a switching state of the converter in the format [P haseA P haseB P haseC ].
The triangles can be divided into lower triangles and upper triangles depending
on their orientation and their switching sequence, as shown in Fig. 4.4 where the
4 4
lower triangle DEF is marked with "L" and the upper triangle DEF is marked
with "U". During each switching period Tsw , the triangle within which the voltage
vector is located, as defined by the α0 β 0 -coordinates in (4.11), becomes active.

442
331
β' 220

443
332 432
F G 321
221
110 U 210
υ*
444 L
333 433 422
222 322 311
111 D E a' 200
211
000 100

Fig. 4.4: First sector of regions 0 and I indicating the switching states of the lower and
upper triangles

63
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

Then, its switching states vary in the sequence which is marked with red arrows in
Fig. 4.4. More specifically, for the upper triangles the switching sequence will be
t1 TF/2 TG/2 t2 t2 TG/2 TF/2 t1

E F G E E G F E
Tsw
V'E V'F V'G V'E V'E V'G V'F V'E
+[1 1 1] +[1 1 1] (4.14)

and for the lower triangles


t1 TE/2 TF/2 t2 t2 TF/2 TE/2 t1

D E F D D F E D
Tsw
V'D V'E V'F V'D V'D V'F V'E V'D
+[1 1 1] +[1 1 1] (4.15)

SVM - Version 1:
There are now two ways to continue with the generation of the PWM pulses for
the switches of the N -level converter. The first way is to calculate the the dwell
times of each switching state (which is the percentage during the switching period
when the specific state is active). For the upper triangle, the dwell times are given
by [127]
TE = ceil(vβ∗ 0 ) − vβ∗ 0 Tsw


TF = ceil(vα∗ 0 ) − vα∗ 0 Tsw (4.16)




TG = Tsw − TE − TF
and for the lower triangles they are
TE = vα∗ 0 − f loor(vα∗ 0 ) Tsw


TF = vβ∗ 0 − f loor(vβ∗ 0 ) Tsw (4.17)




TD = Tsw − TE − TF
ceil(x) and f loor(x) are mathematical operators for rounding towards up and
down, respectively. The dwell times t1 and t2 in (4.14) and (4.15) are distributed
among the redundant states VE0 for the upper and VD0 for the lower triangles
t1 + t2 = TE (for upper triangles)
(4.18)
t1 + t2 = TD (for lower triangles)
The standard method is to distribute the dwell times t1 and t2 equally as follows
TE
t1 = t2 = (for upper triangles)
4 (4.19)
TD
t1 = t2 = (for lower triangles)
4

64
4.3. Multilevel PWM Modulation Strategies

However, different distributions of the redundant states dwell times is proposed in


the following sections.
Since now the dwell times for each switching state of the converter is known, the
final step is to translate these switching states into pulses for each switch of the
converter. Table 4.2 can be used for that in the case of the 5-level NPC converter.
The main disadvantage of this SVM method is that the final PWM pulses of the
switches are calculated directly from the modulation algorithm, which is difficult
to be implemented on a standard MCU board. Therefore, this version of the SVM
algorithm can be used mainly for computer simulations.

SVM - Version 2:
Another version of the N-level SVM is to calculate the duty cycle of the switches
instead of the dwell times of each voltage vector. The duty cycles for the switching
states of the upper triangles are

dE = ceil(vβ∗ 0 ) − vβ∗ 0
dF = ceil(vα∗ 0 ) − vα∗ 0 (4.20)
dG = 1 − dE − dF
d1 dF dG d2

E F G E
d1+d2=dE (4.21)
and for the lower triangles

dE = vα∗ 0 − f loor(vα∗ 0 )
dF = vβ∗ 0 − f loor(vβ∗ 0 ) (4.22)
dD = 1 − dE − dF
d1 dE dF d2

D E F D
d1+d2=dD (4.23)
Similarly to SVM - version 1, the standard distribution of the duty cycles d1 and
d2 is the following
dE
d1 = d2 = (for upper triangles)
2 (4.24)
dD
d1 = d2 = (for lower triangles)
2
Different distributions of the duty cycles of these redundant states will be proposed
in the following sections.

65
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

By using (4.12) and (4.20)-(4.23) the following matrices can be generated for the
duty cycle of each converter state
 
dP haseA State1 dP haseB State1 dP haseC State1
 dP haseA State2 dP haseB State1 dP haseC
 
State1 
 
dstate =  dP haseA State3
 dP haseB State1 dP haseC State1


 ... ... ...
 

dP haseA State(N −1) dP haseB State(N −1) dP haseC State(N −1)
(4.25)
and for the duty cycle of each switch for the upper-leg of an N -level NPC converter
 
dP haseA switch1 dP haseB switch1 dP haseC switch1
 dP haseA switch2 dP haseB switch2 dP haseC switch2
 

 
dsw =  dP haseA switch3
 dP haseB switch3 dP haseC switch3 

 ... ... ...
 

dP haseA switch(N −1) dP haseB switch(N −1) dP haseC switch(N −1)
(4.26)
The switches of the lower-leg of each converter phase are complimentary to the
upper switches. The detailed code for generating (4.25) and (4.26) is included in
Appendix A.
The duty cycle dsw of (4.26) is sent to the ePWM module of the MCU, as shown
in Fig. 4.5, which generates the corresponding PWM signal by comparing the duty
cycle with a triangular carrier. For a 3-phase N-level NPC converter 3 · (N − 1)
ePWM modules are needed.

MCU
ePWM1
Vˆph* dsw (1,1)
+
-
ePWM2 Gate
SVM dsw (1,2) Driver
+
θ -

Fig. 4.5: Block diagram of the N -level SVM and the ePWM module of the MCU

66
4.4. Electric Drive Design using 3-level NPC Converters

4.4 Electric Drive Design using 3-level NPC


Converters
An electric drive design with 3-level NPC converters is described in this section
and the schematic diagram of the system is shown in Fig. 4.6. The model
and the simulation results of the power converters have been implemented on
Matlab/Simulink and PLECS.
The power switches of the converters are the SiC MOSFETs SCH2080KE with
breakdown voltage VDSmax = 1.2 kV and the clamping diodes are the SiC schottky
diodes C3D10170H with peak reverse voltage VRRM = 1.7 kV. In order to be able
to operate the VSC at the required voltage level, two MOSFETs connected in
series are used for each power-switch-block and two series-connected clamping
diodes for each diode-block of the converter. SiC power switches have been
chosen for the simulation analysis of this study, because they have higher voltage
capabilities compared to Si switches and also lower losses. The losses of the
MOSFETs SCH2080KE and the diodes C3D10170H are modelled according to the
information given in their datasheets. The model on PLECS of the switching and
conduction losses for the SCH2080KE can be seen in Fig. 4.7.
The following figures present the simulation results of the electric drive. Three
operating points of the generator are chosen in order to test the steady-state drive
performance at low, medium and high power, as shown in Fig. 4.8(a). These
three operating points are at 500 rpm / 1.5 kW, at 2000 rpm / 10.5 kW and at
4000 rpm / 46 kW. The dc-link voltages and the generator current and voltage at
the third operating point are shown in Fig. 4.8(c)-(d).

Generator DC-link Phase c Grid


Phase b Phase b

Converter Phase a Phase a Converter


T1
T1
C2
VC2
Generator T2 D1a Vdc tot D1a T2
PCC
Tm (+) (+) Grid
D1b
0
D1b
Grid ZS
PMSG T3 D2a T3 Filter
D2a
C1
D2b VC1 D2b
igen vgen T4 T4 is vs

θgen Vdc tot is


Generator Grid
ωgen Control
SVM SVM vs
Vref Control
DC link
control Pref Qref

Fig. 4.6: Schematic of 3-level generator drive

67
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

(a) (b)

(c)
Fig. 4.7: (a) Turn-on losses, (b) turn-off losses and (c) conduction losses of the MOSFET
SCH2080KE, as modelled on PLECS

The harmonics of the PMSG terminal voltage and current are analyzed in Fig. 4.9,
while the system operates at rated power (operating point 3). The total harmonic
distortion of the PMSG current (T HDIGen ) is equal to 1.8% and the THD of the
phase voltage T HDV phGen is 54.7%.
The efficiency of the two 3-level VSCs is evaluated on PLECS and the simulation
result can be seen in Fig. 4.10. The system efficiency is simulated at these three
operating points and the efficiency plot is extracted through curve-fitting.

68
4.4. Electric Drive Design using 3-level NPC Converters

Total dc-link volt. Vdc tot [V]


Generator Power Pgen [kW]
5000 4200
0
Rotor Speed [rpm]
4000 X: 0.2995
Y: -1.479
4100
X: 0.7431 ngen [rpm]
3000 Y: -11.13 n*gen [rpm] -20 4000
2000 Pgen [kW]
-40 3900
1000
X: 1.418
Y: -46.19
0 -60 3800
0 0.5 1 1.5 0.5 1 1.5
Time [s] Time [s]
(a) (b)

Phase voltage - VgenA [V]


2010 20
2000
DC-link voltage [V]

2005 10
1000

IgenA [A]
2000 0 0

-1000
1995 VC1 Vgen a -10
VC2 -2000 Igen a
1990 -20
1.45 1.46 1.47 1.48 1.49 1.5 1.47 1.471 1.472 1.473 1.474
Time [s] Time [s]
(c) (d)

Fig. 4.8: Simulation results of the 3-level generator drive: (a) rotor speed and generator
power, (b) total dc-link voltage, (c) dc-link capacitor voltages and (d) phase
voltage and current at the generator-side converter
Generator Current IgenA [A]
VgenA [V]

3-level generator-side VSC, THDVphGen=54.7%


0.5
3-level generator-side VSC, THDIGen=1.8%
2000
0.4
1500
0.3
Phase Voltage

1000
0.2
500 0.1

0 0
0 10 20 30 40 0 10 20 30 40
Frequency [kHz] Frequency [kHz]
(a) (b)

Fig. 4.9: Simulation results of the 3-level generator drive at rated power (4000rpm/46kW):
(a) FFT of the PMSG phase voltage and (b) current

69
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

Converters' Efficiency [%]


99 8
X: 46.17 X: 1.535
X: 11.24
Y: 98.74

THDIGen [%]
Y: 98.61 6 Y: 5.405
98
X: 10.99
4 Y: 3.287
X: 46.93
97 Y: 1.81
X: 1.479 2
Y: 96.81
96 0
0 10 20 30 40 10 20 30 40
Pgen [kW] Pgen [kW]

(a) (b)

Fig. 4.10: Simulation results of the 3-level (a) power converters’ efficiency and (b) THD
of the stator current

4.5 Electric Drive Design using 5-level NPC


Converters
The electric drive design with 5-level NPC converters is presented in this section.
The same power switches and diodes with the ones mentioned in the previous
section are used for this drive, namely the 1.2 kV SCH2080KE SiC MOSFETs and
the 1.7 kV C3D10170H SiC diodes.
The main challenge in the control of the 5-level NPC converter is the fact that
the converter is naturally unbalanced at the dc-side and, therefore, without any
special control of its four dc-link capacitors, their voltage will variate a lot from
its set-point [Fig. 4.11(a)]. As can be seen on the right side of Fig. 4.11(b),
the unbalancing of the dc-link capacitor voltages can create distortions on the
ac voltage of the converter and additional harmonics on the ac current. It can
also cause the capacitors of the dc-link and the power switches of the converter to
exceed their maximum voltage capabilities.
The available methods for balancing the dc-link capacitor voltages in a drive
with 5-level NPC converters are divided into two types, the software and the
hardware based methods. Many control techniques have been developed based
on a modified version of the SVM by utilizing the redundant states of the 5-level
space vector plane. The control of the redundant states in the SVM can affect the
dc-link capacitor voltages and at the same time, the requested line-to-line voltage
at the ac side is created. However, the effectiveness of purely software based
control techniques is limited, since they cannot provide voltage balancing when
the converter is operating with resistive power factor and high modulation index,
as described in [129]. Therefore, some modifications and operational limitations

70
4.5. Electric Drive Design using 5-level NPC Converters

2
1
1.5 0.5

VLL [pu]
Vdc [pu]

Vc1
1 0
Vc2 Vab
0.5 Vc3 -0.5 Vbc
Vc4 -1 Vca
0
0.1 0.12 0.14 0.16 0.18 0.1 0.12 0.14 0.16 0.18
Time [s] Time [s]
(a) (b)

Fig. 4.11: 5-Level NPC converter without dc-link balancing control: (a) the four dc-link
capacitor voltages and (b) the L2L output voltages

should be applied on the NPC converter. Specifically, a voltage balancing method


in a 5-level NPC converter used in active filters has been proposed in [72]. Similar
modulation techniques have been proposed in [73, 74]; however, special design of
the passive filter is required in [74] and the balanced operation can be accomplished
only with capacitive or inductive power factor in [73]. An SVM-based control
strategy for HVDC systems with two B2B-connected 5-level NPC converters
has been presented in [75], where the capacitor voltages are balanced effectively.
However, the modulation index of both NPC converters has to be the same and
this is not possible in an electric drive system with a variable speed generator.
Hardware based voltage balancing techniques usually employ multistage dc/dc
converters at the dc-link. These dc/dc converters are bidirectional buck-boost
converters that control usually the voltage of the four capacitors in pairs. Various
control techniques have been developed, such as in [83–85], which can provide
effective voltage balancing in every operational condition of the electric drive.
However, their main disadvantages are the increased installation cost and size
of the system due to the additional hardware, as well as the reduced system
efficiency because of the losses on the dc/dc converter. A balancing converter with
reduced number of passive components was proposed in [86], but it requires more
power switches and the control is more complicated than the previously mentioned
topologies.
This sections presents three alternative ways of balancing the dc-link capacitor
voltages:
• A purely software-based balancing control is presented in Subsection
4.5.1, where a new modified SVM algorithm is used for controlling the four
capacitor voltages. This voltage control scheme is based on the quasi-3-level

71
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

control proposed for single NPC converters in [82].


• Hardware-based voltage balancing is shown in Subsection 4.5.2 where
different types of control for the balancing converter are presented.
• A combination of the hardware and software-based balancing is an-
alyzed in Subsection 4.5.3, which proposes a balancing converter with reduced
number of components, utilizing also the SVM for balancing the voltages
under certain operating conditions.

2
4.5.1 SVM-based DC-Link Voltage Balancing
The multilevel SVM has got redundant states at many points of the space vector
plane. These redundant states create the same L2L voltage at the ac side of the
converter. However, each state has different effect on the capacitor voltages of the
dc-side, which can be utilized to balance their voltages. The 5-level SVM diagram
of Fig. 4.3 can be divided into three regions depending on the modulation index
m̂:
• Region 0 (0 ≤ m̂ < 0.25): The modulation index at this region is low and
the 5-level NPC converter generates 2-level L2L voltage at the ac side. There
are four different redundant states that can control the voltage level of the
four dc-link capacitors.
• Region I (0.25 ≤ m̂ < 0.5): The 5-level NPC converter has 3-level L2L
voltage at the ac side. As can be seen in Fig. 4.4, there are four different
redundant states to be chosen for controlling the switches.
• Region II (0.5 ≤ m̂ ≤ 1): The converter has 4-level voltage between 0.5 ≤
m̂ < 0.75 and 5-level voltage for m̂ ≥ 0.75. The redundant states in this
region are not enough to ensure balanced condition with every load of the
converter [129]. Therefore, the switching sequence of the SVM will be
modified, as described later in this section.
The structure of an generator drive with two 5-level NPC converters and a PMSG is
illustrated in Fig. 4.12. In this subsection a fully software-based voltage balancing
technique is shown and the first step is to split the dc-link capacitors between the
two converters, as shown in Fig. 4.12. The NPC converters are connected to each
other only at the top and the bottom joint-points (points 5 and 1, respectively)
2 More details can be found:

G. Mademlis and Y. Liu, "DC Link Voltage Balancing Technique Utilizing Space Vector Control
in SiC-based Five-Level Back-to-Back-Connected NPC Converters", in Proc. of the IEEE Energy
Conversion Congress and Exposition (ECCE), September 23-27, 2018.

72
4.5. Electric Drive Design using 5-level NPC Converters

Phase c Phase c
Generator Phase b DC-link Phase b
Grid
Converter
Id 5a Id 5b Converter
Phase a Phase a
5.
T1 T1
C4b C4a
T2 D1 T2
D1 Id 4a Id 4b
4.
T3 T3
D3 D3
Generator C3a
C3b PCC
Tm (+) T4
Id 3a 3. Id 3b T4 (+) Grid
D5 D5 Grid ZS
PMSG Filter
T5 D2 D2 T5
C2b C2a is vs
igen vgen T6 T6
D4 Id 2a Id 2b D4
2.
T7 T7
D6 C1b C1a D6
igen vgen
T8 Id 1a 1. Id 1b T8

θgen Vdc tot is


Generator VC 1÷4 b VC 1÷4 a Grid
ωgen Control
SVM SVM vs
Vref Control
DC link
control
Pref Qref

Fig. 4.12: Structure of a generator drive with two 5-level NPC converters connected B2B
using modified SVM for voltage balancing of the dc-link

of the dc-link. The parallel connection of the two groups of capacitors (Ca and
Cb ) will ensure that they will have the same voltage. The middle point of the
two converters can be connected to the ground or the neutral point of the system.
The two converters operate at different modulation indexes and a common dc-link
capacitor bank would have caused interferences that would have affected the
balancing of the dc-link voltages. The configuration presented here decouples
the generator-side converter of the drive from the grid-side converter allowing
individual voltage balancing of the two groups Ca and Cb of the dc capacitors. The
total dc-link voltage of the drive Vdc tot is still controlled by the dc-link controller
of the grid converter (Subsection 2.3.1).
The SVM diagram for a 5-level converter can be seen in Fig. 4.13(a) and the
modulation strategy described in this section is identical for controlling both
NPC converters of the electrical machine drive. The grid-side converter operates
at an almost constant modulation index m̂grid ≤ 1 in region II, whereas, the
generator-side converter has variable modulation index m̂gen , since the speed and
the voltage of the generator are changing. The variable v ∗ in Fig. 4.13 is the
normalized reference voltage of the converter, as calculated by (4.10), and for a
5-level converter it varies from 0 to 4.

Voltage balancing algorithm for regions 0 & I: The switching states and
the switching sequence of the SVM on the triangles of the space vector plane
are indicated with arrows in Fig. 4.13(b). When the NPC converter operates

73
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

ib -ic
2nd Sector 442
331
040 140 240 340 440
θο Regions 0 & I: 220

141 241 341 441


3rd 041
030 130 230 330
430 1st 443
432
332 F G
Sector 242 342 442
Sector 221 321
042
142
131 231 331
431
420 110 210
031
020 120 220
320 U
343 443
043
143
243
132
232 332
432
321
421
410
444
333
L υ* 422
032
021
121 221
210
310 υ*
010 444 110
θ
222 311
-ia 144
244
344
233
333
433
322
422
411 ia 111 D E 200
044 133 222 311 400
033
022
122
111
211
200
300 000
011
Region
334 0
000 434
100 F
234 423 Region II: 420
134 223 323 412
034 123 312 401
023 112 212 301
012 201
001 101
224
Region
324
I 424
124 413
024 113 213 313 402
013 302
4th 002 102 202 6th 2∙ 210 421
410
Sector 014
114 214 314 414
403
Sector 310
υ*
003 103 203 303
Region II υ*
004 104 204 304 404 422
th 311 400
5 Sector
200 D E
ic -ib 2∙211 2∙200
2∙100

(a) (b)

Fig. 4.13: Five-level SVM diagram: (a) space vector plane indicating the regions and
sectors; (b) first sector of of the SVM plane indicating the switching sequence

in regions 0 and I, there are four available redundant switching states in each
switching cycle that can easily be assigned to each of the four dc-link capacitors
in order to charge them or discharge them whenever it is necessary.
An example is shown in Fig. 4.14, where the effect of the redundant states on the
state of charge of the dc-link capacitors is demonstrated. The circuit diagram of the
NPC converter is shown there, when the redundant states {100}/{211}/{322}/{433}
from Sector 1 of the SVM plane are active. The switches which are active in each
case are shown with thick black lines, while the inactive switches are marked with
grey line. Specifically, when the instantaneous phase A current ia is positive in Fig.
4.14(a), the state {100} discharges the capacitor C1 . Oppositely, if the current
ia is negative, the capacitor C1 gets charged. Similarly, in Fig. 4.14(b) the state
{211} discharges the capacitor C2 when ia > 0. In Fig. 4.14(c) the state {322}
discharges the capacitor C3 and in Fig. 4.14(d) the state {433} discharges the
capacitor C4 . The L2L voltage at the ac side of the NPC converter is the same in
all the four cases.
The three-phase current at the ac side of the NPC converters and the four capacitor
voltages are measured and compared with each other in order to choose the correct
redundant states from the space vector diagram. The sign of the ac current affects
also the charging state of the dc-link capacitors, as has been shown in Fig. 4.14.
The current that affects the capacitor voltages at each time instant depends on the

74
4.5. Electric Drive Design using 5-level NPC Converters

C4 C4

TA3
C3 TA4 C3 TA4 TB4 TC4
0 0
TA5 TB5 TC5 iC2 TA5 TB5 TC5
C2 TA6 TC6
C2 TA6 TC6
TB6 TB6
iC1 TA7 TC7 TB7 TC7
TB7
C1 ia TB8 TC8 C1 ia ib ic
ib ic

a b c a b c
(a) (b)

iC4 TA1
C4 TA2 C4 TA2 TB2 TC2
iC3 TA3 TB3 TC3 TA3 TB3 TC3
C3 TA4 TB4 TC4
C3 TA4 TB4 TC4
0 0
TA5 TB5 TC5 TB5 TC5
C2 TB6
C2
TC6

C1 ia ib ic C1 ia ib ic

a b c a b c
(c) (d)

Fig. 4.14: Circuit diagram of the 5-level NPC converter when the switching state (a)
{100}, (b) {211}, (c) {322} and (d) {433} is active

Table 4.4: Balancing Current [78]

Voltage angle θ Balanc. current ibl


(-π/6, π/6] ia
(π/6, π/2] −ic
(π/2, 5π/6] ib
(5π/6, 7π/6] −ia
(7π/6, 3π/2] ic
(3π/2, 11π/6] −ib

75
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

Table 4.5: Voltage balancing algorithm for region 0 (lower triangles)

ibl Condition Switching sequence t1 t2


>0 VCmax = VC1 {VD ; VE ; VF ; VD + [1 1 1]} TD /4 TD /4
>0 VCmax = VC2 {VD + [1 1 1]; VE + [1 1 1]; VF + [1 1 1]; VD + [2 2 2]} TD /4 TD /4
>0 VCmax = VC3 {VD + [2 2 2]; VE + [2 2 2]; VF + [2 2 2]; VD + [3 3 3]} TD /4 TD /4
>0 VCmax = VC4 {VD + [3 3 3]; VE + [3 3 3]; VF + [3 3 3]; VD + [4 4 4]} TD /4 TD /4
<0 VCmin = VC1 {VD ; VE ; VF ; VD + [1 1 1]} TD /4 TD /4
<0 VCmin = VC2 {VD + [1 1 1]; VE + [1 1 1]; VF + [1 1 1]; VD + [2 2 2]} TD /4 TD /4
<0 VCmin = VC3 {VD + [2 2 2]; VE + [2 2 2]; VF + [2 2 2]; VD + [3 3 3]} TD /4 TD /4
<0 VCmin = VC4 {VD + [3 3 3]; VE + [3 3 3]; VF + [3 3 3]; VD + [4 4 4]} TD /4 TD /4

VCmax , VCmin are the maximum and minimum capacitor voltage, respectively

Table 4.6: Voltage balancing algorithm for region I (lower triangles) [72]

ibl Condition 1 Condition 2 Switching sequence t1 t2


>0 VCmax1 = VC1 - {VD ; VE ; VF ; VD + [1 1 1]} TD /3 TD /6
>0 VCmax1 = VC2 VCmax2 = VC1 {VD ; VE ; VF ; VD + [1 1 1]} TD /6 TD /3
>0 VCmax1 = VC2 VCmax2 = VC3 {VD + [1 1 1]; VE + [1 1 1]; VF + [1 1 1]; VD + [2 2 2]} TD /3 TD /6
>0 VCmax1 = VC3 VCmax2 = VC2 {VD + [1 1 1]; VE + [1 1 1]; VF + [1 1 1]; VD + [2 2 2]} TD /6 TD /3
>0 VCmax1 = VC3 VCmax2 = VC4 {VD + [2 2 2]; VE + [2 2 2]; VF + [2 2 2]; VD + [3 3 3]} TD /3 TD /6
>0 VCmax1 = VC4 - {VD + [2 2 2]; VE + [2 2 2]; VF + [2 2 2]; VD + [3 3 3]} TD /6 TD /3
<0 VCmin1 = VC1 - {VD ; VE ; VF ; VD + [1 1 1]} TD /3 TD /6
<0 VCmin1 = VC2 VCmin2 = VC1 {VD ; VE ; VF ; VD + [1 1 1]} TD /6 TD /3
<0 VCmin1 = VC2 VCmin2 = VC3 {VD + [1 1 1]; VE + [1 1 1]; VF + [1 1 1]; VD + [2 2 2]} TD /3 TD /6
<0 VCmin1 = VC3 VCmin2 = VC2 {VD + [1 1 1]; VE + [1 1 1]; VF + [1 1 1]; VD + [2 2 2]} TD /6 TD /3
<0 VCmin1 = VC3 VCmin2 = VC4 {VD + [2 2 2]; VE + [2 2 2]; VF + [2 2 2]; VD + [3 3 3]} TD /3 TD /6
<0 VCmin1 = VC4 - {VD + [2 2 2]; VE + [2 2 2]; VF + [2 2 2]; VD + [3 3 3]} TD /6 TD /3

VCmax1 , VCmax2 are the highest and the second highest capacitor voltage.

VCmin1 , VCmin2 are the lowest and the second lowest capacitor voltage.

angle θ of the reference voltage v ∗ and the proper balancing current ibl is chosen
according to Table 4.4, as referenced from [78].
It was mentioned in (4.18) that the dwell times t1 and t2 are distributed among the
redundant states VE0 for the upper and VD0 for the lower triangles. The standard
way is to distribute them evenly, in order to minimize the switching losses of the
converter and this method from now on will be called as "standard SVM" [127].
However, it is possible by varying these variables to control the capacitor voltages,
as shown in Table 4.5 for the lower triangles of region 0 and in Table 4.6 for region

76
4.5. Electric Drive Design using 5-level NPC Converters

I (which is referenced from [72]). Similar methodology is followed for deriving the
switching states for the upper triangles.

Voltage balancing algorithm for region II: The redundant states in the
triangles of region II are not enough to keep all the capacitor voltages balanced,
as can be seen in Fig. 4.13(b). More specifically, there are only two available
redundant states on the D edge of the outer lower triangles in region II, the {300}
and {411}, and three redundant states on the D edge of the inner lower triangles,
the {200}, {311} and {422}.
One way to overcome this issue is to treat the 5-level NPC converter as if it was
a 3-level converter. In this case, the capacitor voltages are controlled in pairs,
VC1 + VC2 and VC3 + VC4 , and the redundant states are enough to balance them
effectively. This control strategy has been proposed in [72, 82] for single NPC
converters and it is applied here for a drive with two B2B NPC converters. The
switching states that generate current at the clamping points 2 and 4 of the NPC
converters (Fig. 4.12) are not used in this SVM scheme (such as the states {300},
{311} and {411}) and the currents Id2a , Id4a , Id2b and Id4b are zero. Therefore,
there is no voltage drift between the capacitor voltages of each pair. The resultant
redundant states marked with a red box in Fig. 4.13(b) are utilized to keep the
average current of the middle point 3 equal to zero, in order to balance the two
pairs of the capacitor voltages. A more detailed view of the effect of these switching
states on the charging state of the dc-link capacitors and the dc-side currents can
be seen in Fig. 4.15.
The conversion from 5-level operation into 3-level operation is implemented by
converting the large triangle on the bottom side of Fig. 4.13(b) into the equivalent
triangle on the left side. Therefore, the reference voltage vector of the SVM in
(4.10) is now modified as follows

0 v∗ 3V̂ph
v∗ = = (4.27)
2 4Vdc
0
The modified reference voltage v ∗ is now used in the rest of the equations (4.11)-
(4.13) in order to calculate the space vectors and the dwell times. It can be seen in
Fig. 4.13(b) that there are two available redundant states in each triangle, which
are selected as shown in Table 4.7.

77
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

id5 TA1
C4 TA2 C4 TA2
id4
TA3 TA3
C3 TA4 C3 TA4 TB4 TC4
0 0
TA5 TB5 TC5 TB5 TC5
C2 TC6
C2 TC6
TB6 TB6
id2
TB7 TC7 TB7 TC7
C1 ia TB8 TC8 C1 ia ib ic
ib ic

a b c a b c
(a) (b)

TA1
C4 C4 TA2

TA3 iC3 + iC4 TA3 TB3 TC3


C3 TA4 C3 TA4 TB4
id3 TC4
0 0
TA5 TB5 TC5 TB5 TC5
C2 TA6 TC6
C2 TB6
TB6 TC6
iC1 + iC2
TB7 TC7
C1 ia TB8 TC8 C1 ia ib ic
ib ic

a b c a b c
(c) (d)

Fig. 4.15: Circuit diagram of the 5-level NPC converter when the following switching
states are active: (a) {300}, (b) {411}, (c) {200} (⇒ the C1 and C2 are getting
discharged) and (d) {422} (⇒ the C3 and C4 are getting discharged)

Table 4.7: Voltage balancing algorithm for region II (lower triangles)

ibl Condition Switching sequence t1 t2


>0 VC1 + VC2 > VC3 + VC4 2·{VD ; VE ; VF ; VD + [1 1 1]} TD /3 TD /6
>0 VC1 + VC2 < VC3 + VC4 2·{VD ; VE ; VF ; VD + [1 1 1]} TD /6 TD /3
>0 VC1 + VC2 > VC3 + VC4 2·{VD ; VE ; VF ; VD + [1 1 1]} TD /6 TD /3
>0 VC1 + VC2 < VC3 + VC4 2·{VD ; VE ; VF ; VD + [1 1 1]} TD /3 TD /6

Simulation results: The generator drive of Fig. 4.12 has been simulated on
Matlab/Simulink. The parameters of the system are the same with the ones
for the 3-level drive reported in Table 4.1. The grid-side converter operates at

78
4.5. Electric Drive Design using 5-level NPC Converters

1002
VC1a

Voltage [V]
VC2a
1000 VC3a
VC4a
998
0.27 0.275 0.28 0.285 0.29 0.295 0.3
Time [s]
1002
VC1b
Voltage [V]

VC2b
1000 VC3b
VC4b
998
0.27 0.275 0.28 0.285 0.29 0.295 0.3
Time [s]

(a)

1002
VC1a
Voltage [V]

VC2a
1000 VC3a
VC4a
998
0.5 0.505 0.51 0.515 0.52 0.525 0.53 0.535 0.54 0.545 0.55
Time [s]
1002
Voltage [V]

VC1b
VC2b
1000
VC3b
VC4b
998
0.5 0.505 0.51 0.515 0.52 0.525 0.53 0.535 0.54 0.545 0.55
Time [s]

(b)

1002
Voltage [V]

VC1a
VC2a
1000
VC3a
VC4a
998
1.05 1.055 1.06 1.065 1.07 1.075 1.08 1.085 1.09 1.095 1.1
Time [s]
1002
Voltage [V]

VC1b
VC2b
1000
VC3b
VC4b
998
1.05 1.055 1.06 1.065 1.07 1.075 1.08 1.085 1.09 1.095 1.1
Time [s]

(c)
Fig. 4.16: Simulation results of the dc-link voltages when the generator-side 5-level NPC
converter operates in (a) region 0, (b) region I and (c) region II of the SVM
plane

79
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

constant modulation index which is at region II of the SVM diagram. However,


many operating points of the generator are simulated in order to investigate all
the three regions of the SVM for the generator-side converter. The same three
operating points shown in Fig. 4.8(a) (500 rpm / 1.5 kW operating at region 0,
2000 rpm / 10.5 kW at region I and 4000 rpm / 46 kW at region II) have been
simulated also here.
The reference voltage of each capacitor bank is equal to 4000/4 = 1000 V. As can
be seen in Fig. 4.16, the capacitor voltages remain balanced in all the operating
regions of the system (the blue curve of VC1a is overlapped by the red one of
VC2a and the black curve of VC3a is overlapped by the green curve of VC4a ). The
Vc1a ÷ Vc4a are the capacitor voltages of the grid-side NPC converter and the
Vc1b ÷ Vc4b are at the generator-side. The voltages in Fig. 4.16(a) are slightly
declining because the converter has not fully reached steady state yet.
The L2L voltage of the generator terminals can be seen in Fig. 4.17, where the
2-level voltage of region 0 is in Fig. 4.17(a) and the 3-level voltage of regions I
and II is in Figs. 4.17(b)-(c). It can be seen here that although the generator-side
converter operates with high modulation index in region II, the voltage still has 3

4
4
2
Vgen AB [kV]

Vgen AB [kV]

2
0
0
-2 -2

-4 -4
0.27 0.275 0.28 0.285 0.29 0.295 0.3 0.5 0.505 0.51 0.515 0.52 0.525 0.53
Time [s] Time [s]
(a) (b)
4

2
Vgen AB [kV]

-2

-4
1.05 1.055 1.06 1.065 1.07 1.075 1.08
Time [s]
(c)
Fig. 4.17: Simulation results of the PMSG L2L voltage when the generator-side 5-level
NPC converter operates in (a) region 0, (b) region I and (c) region II of the
SVM plane

80
4.5. Electric Drive Design using 5-level NPC Converters

0.99

Converters' Efficiency [%]


X: 46.39
X: 10.96
Y: 0.9877
Y: 0.9865
0.98

0.97
X: 1.246
Y: 0.9592
0.96
0 10 20 30 40
Pgen [kW]

Fig. 4.18: Simulation results of the 5-level power converters’ efficiency when SVM-based
dc-link voltage balancing is used

levels, due to the voltage balancing algorithm described previously. This is also
the main disadvantage of the SVM-based voltage balancing method, because the
converters are not possible to generate 5-level voltage, although the voltage stress
on each MOSFET is still equal to Vdc tot /4.

3
4.5.2 Hardware-based DC-Link Voltage Balancing
The dc-link voltage balancing of NPC-based electric drives can be implemented
with specially designed dc/dc converters placed in parallel to the dc-link, as
described in [84, 88, 130, 131] and shown in Fig. 4.19. The dc/dc converters can be
controlled with constant duty cycle in open-loop mode or they can have a closed
loop control system. Both ways are described in this subsection.

Open-loop Control of the Balancing Converter

The dc/dc converter balances the capacitor voltages in pairs and, therefore, three
stages are needed for a 5-level drive. Each converter stage consists of two power
switches with antiparallel diode and an inductor. The maximum instantaneous
voltage applied across each switch is equal to

vTbk = vCk + vC(k+1) (4.28)

which is two times the rated voltage of the switches of the NPC converter. As
can been seen in Fig. 4.19, the dc/dc converter has been modelled here with
two series-connected switches in each switch position. Instantaneous currents and
3 More details can be found:

G. Mademlis and Y. Liu, "Feed-forward Control of Active Voltage Balancing Converter in Electric
Drive with Five-Level NPC Converters", in Proc. of the 2018 IEEE Energy Conversion Congress
and Exposition (ECCE), September 23-27, 2018.

81
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

DC-link Balancing Converter


Id 5a Id 5b
Stage 3

VC4 Tb3'
C4
Id 4aStage 2 Id 4b
L3
Three-phase Tb2 C3 VC3 Tb3 Grid
Three-phase
PMSG 5-level NPC Id 3a Id 3b Grid
5-level NPC
Generator- Filter
Grid-side
side L2
Tb2' C2 VC2 Tb1
Id 2a Id 2b
L1
C1 VC1 Tb1'
Id 1a Id 1b
Stage 1
VC 1÷4

Fig. 4.19: Power circuit of the the dc/dc converter that balances the dc-link of a 5-level
drive when only hardware balancing is used

voltages are symbolized here with lowercase letters (i and v respectively), while
average quantities are shown with capital letters (I and V ).
The two switches Tbk and Tbk0 are complimentary to each other and the duty
cycle of the k-stage power switch is defined as [132]
vC(k+1)
d0k = 1 − dk = (4.29)
vCk + vC(k+1)

When the capacitor voltages of that converter stage are equal


(4.29)
vCk = vC(k+1) ⇒ dk = d0k = 0.5 (4.30)

Therefore, by applying a constant duty cycle equal to 0.5 on all the power switches,
the capacitor voltages will have the same voltage.
The operation of the converter can be explained by the graphs in Fig. 4.20. The
first stage of the balancing converter is shown as an example and the other two
stages operate similarly:
• When VC1 > VC2 [Fig. 4.20(a)]
1. the capacitor C1 charges the inductor and the current flows through
the switch Tb1
0
during half of the switching period.

82
4.5. Electric Drive Design using 5-level NPC Converters

2. During the other half period, the switch Tb1 0


is off and the Tb1 is on.
However, the current does not pass through the switch Tb1 but through
its antiparallel diode and the energy of the inductor is finally transferred
to the capacitor C2 .
• When VC1 < VC2 [Fig. 4.20(b)]
1. the capacitor C2 charges the inductor and the current flows through
the switch Tb1 during half of the switching period.
2. During the other half period, the switch Tb1 is off and the Tb1
0
is on.
The current now passes through the antiparallel diode of Tb1 and the
0

energy of the inductor is transferred to the capacitor C1 .


The average currents of the switches in Fig. 4.20 are defined as

IT B1 = IL1 db1
(4.31)
IT B10 = IL1 (1 − db1 )

The following current and voltage equations exist in the circuit of Fig. 4.20

iT B10 = iC1 + id1a − id1b (4.32)


vT B1 = vC2 − vL1 (4.33)

Since the antiparallel diodes of the switches conduct almost half of the time, it
is recommended to use power MOSFETs for this converter, due to their ability
to have reverse conduction current [133]. MOSFETs allow positive, as well as
negative current to flow through their N -channel, when the gate-source voltage
is positive. The N -channel of the MOSFET has lower conduction and switching

VC 1 > VC 2 VC 1 < VC 2

IC 2
IC 2
C2 Tb1 C2 1
2 Tb1
Id 2a ITb1 I Id 2a ITb1 Id 2b
d 2b

IL 1 L1 IL 1 L1
C1 Tb1' C1 Tb1'
1 2
Id 1a IC 1 ITb1' Id 1b Id 1a IC 1 ITb1' Id 1b

(a) (b)

Fig. 4.20: Power circuit of the first stage of the dc-link balancing converter when (a) the
capacitor C2 and (b) when C1 are being charged

83
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

losses compared to the antiparallel diode or the body diode. Therefore, by forcing
the current to flow always through the switch even when it is negative, the losses
of the balancing converter can be reduced.
It is very important to define a large enough dead-time between the two compli-
mentary switches of each stage of the balancing converter. Otherwise, there is the
danger of short-circuiting the dc-link capacitors and creating large current spikes
which could destroy the switches.
The open-loop control of the balancing converter offers a simple and effective way
of balancing the dc-link capacitor voltages of the 5-level drive. However, since
there is no feedback of the actual voltage of the capacitors, there will be steady
state error between the four voltages due to the losses and the voltage drop of
the system. The most serious drawback of this control method is that there is no
active control of the current that flows through the dc inductors and it is actually
the resultant of the voltage difference between the capacitors. Therefore, in the
case of a step-change of the load of the NPC converters, high current overshoots
and oscillations are expected on the inductors, which cannot be suppressed, as
seen in Fig. 4.21.

190 50
Inductor current IL3 [A]
Capacitor Voltage [V]

180
0
VC4 X: 0.4907
170 Y: -2.8
VC3
-50
160 VC2
VC1
X: 0.5325
150 -100 Y: -83.2
0.4 0.5 0.6 0.7 0.8 0.4 0.5 0.6 0.7 0.8
Time [s] Time [s]
(a) (b)
10

500
5
Vgen AB [V]

Igen [A]

0 0

-5 I
-500 genq
IgenA
-10
0.4 0.5 0.6 0.7 0.8 0.4 0.5 0.6 0.7 0.8
Time [s] Time [s]
(c) (d)
Fig. 4.21: Experimental results of the 5-level generator drive with a step-change of the
PMSG speed from 450 rpm to 650 rpm when the dc-link is balanced with
open-loop controlled balancing converter and with Vdc tot = 670 V

84
4.5. Electric Drive Design using 5-level NPC Converters

Experimental results of the generator drive, which will be presented in the next
Chapter 5, are shown in Fig. 4.21. Specifically, the total dc-link voltage of the
system Vdc tot is 670 V and a step-change of the generator’s reference speed is
applied at 0.49 s. The current of the NPC converter Igen increases due to the
acceleration of the machine and the sudden current increase creates a larger drift
of the capacitor voltages in Fig. 4.21(a). The current of the dc-inductors increases
as well trying to eliminate the voltage drift and can reach high values as shown in
Fig. 4.21(b) for the IL3 which has peak of 83 A. In that particular case, the IL3
became so high, because the inductor core got saturated.

Feedforward-based Closed-loop Control of the Balancing Converter

Each stage k of the balancing converter is regulated by separate controllers


consisting of two cascaded control loops. The primary control loop is the voltage
controller which creates the current reference that balances the two capacitor
voltages. The secondary control loop is the current controller, which has as output
the duty cycle dk of the balancing converter. The voltage and current control loops
can be made of PI controllers, as described in [84]. A feedforward-based control
has been developed for the specific dc/dc converter in [131, 132] which uses only
proportional controllers and offers faster dynamic response. It has been adapted
in this thesis in order to be applied on the balancing converter of a generator drive
with B2B-connected NPC converters.
The control structure of the dc/dc converter is presented in Fig. 4.22. The PV and
PI are the proportional gains of the voltage and current controllers, respectively.
The control of the first stage of the dc/dc converter can be described as follows.
The reference capacitor currents of each stage IC∗ are given by the error between
the two capacitor voltages, after being amplified by the proportional gain of the
voltage controller PV . The reference current of the switch is calculated by using
(4.6), with feed-forward of the dc-link currents. The reference inductor current is
finally calculated with (4.5). The db1 in (4.5) is the duty cycle of the switch Tb1
during obtained from (4.29) during the previous sample period of the controller.
Input of the current controller is the error between the reference and actual
inductor current. The current error is amplified by the gain PI and becomes the
reference average voltage of the inductor VL∗ . Then, the reference voltage of the
switch VT∗B is calculated by using (4.7). During steady state operation of the
balancing converter, the average voltage of the inductor VL1 is equal to 0 and,

85
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

1st stage
Voltage Controller Current Controller

Id1a V C1
VC2 db1* + PWM
x
+ + V L1* VTb1* ÷÷ Gate Tb1
Ver IC1 * + ITB1'* IL 1* + Ier - +
PV1 x P I1 -
- - - + +
V C1 Id1b 1/(1-db1) VC2 VC1

IL1
nd
2 stage
Voltage Controller Current Controller

Id2a VC2
VC3 db2* + PWM
x
+ + V L2 * UTb2* ÷÷ Gate Tb2
Ver IC2* + ITB2'* IL2* + Ier - +
PV2 x P I2 -
- - - + +
VC2 Id2b 1/(1-db2) VC3 V C2

IL2
3rd stage
Voltage Controller Current Controller

Id5a VC4
VC3 db3* + PWM
x
+ + VL3* UTb3* ÷÷ Gate Tb3
Ver IC4* + ITB3'* IL3* + Ier - +
P V3 x PI3 -
- - - + +
V C4 Id5b 1/(1-db3) VC3 VC4

IL 3

Fig. 4.22: Block diagram of the feed-forward control system of the dc-link balancing
converter

therefore, the (4.7) becomes


VT B1 ≈ VC2 (4.34)

The reference duty cycle of switches in the fist stage of the converter are calculated
with (4.29) and (4.34) as
VC1
d∗b1 ≈ ∗ (4.35)
VT B1 + VC1

For the implementation of the feed-forward control scheme measurements of the


four capacitor voltages, the three inductor currents (IL1 , IL2 and IL3 ) and the
dc-link currents (Id1a up to Id5a and Id1b up to Id5b shown in Fig. 4.12) are
required. The current signals are filtered with LPF in order to suppress high order
harmonics. It is possible to reduce the number of measurements by estimating the
dc currents of the NPC converters. The dc currents are proportional to the ac
output currents and the duty cycle of the NPC converters. A similar study has
been conducted in [85] for single phase 5-level NPC converter. The dc-current
estimation algorithm for a 5-level 3-phase NPC converter can be described as

86
4.5. Electric Drive Design using 5-level NPC Converters

follows for the phase a of the grid-side NPC converter

Id5ba = ISα [dG1α + dG2α + dG3α + dG4α ]


Id4ba = ISα [(1 − dG1α ) + dG2α + dG3α + dG4α ]
Id3ba = ISα [(1 − dG1α ) + (1 − dG2α ) + dG3α + dG4α ] (4.36)
Id2ba = ISα [(1 − dG1α ) + (1 − dG2α ) + (1 − dG3α ) + dG4α ]
Id1ba = ISα [(1 − dG1α ) + (1 − dG2α ) + (1 − dG3α ) + (1 − dG4α )]

where ISα is the grid line current of phase α and dG1α is the duty cycle of the
switch TG1 at the phase α of the grid-side NPC converter. The current estimation
of the other two phases, b and c, can be obtained by similarly utilizing the (4.36).
It should be noted that the currents given by (4.36) are the average values and
not the instantaneous currents. The final values of the dc currents are calculated
by summing the contributions of the three phases

Id5b = Id5ba + Id5bb + Id5bc


Id4b = Id4ba + Id4bb + Id4bc
Id3b = Id3ba + Id3bb + Id3bc (4.37)
Id2b = Id2ba + Id2bb + Id2bc
Id1b = Id1ba + Id1bb + Id1bc

The dc currents of the machine-side NPC converter Id1a up to Id5a are calculated
similarly to equations (4.36) and (4.37) by using the machine-side line currents
Igen instead of the grid current IS in (4.36).
The system shown in Fig. 4.19 has been simulated using Matlab/Simulink and the
simulation results are shown in the following figures. The variation of the dc-link
capacitor voltages is shown in Fig. 4.23(a) and they are balanced with small steady
state error. This verifies the effectiveness of the control system described in the
previous section. The total dc-link voltage is controlled by the grid-side converter
and has mainly remained unaffected by the individual capacitor voltages, as can
be seen in Fig. 4.23(b). The estimated and measured currents at the dc side of
the grid NPC converter are presented in Fig. 4.24. The estimated currents are
calculated by (4.36)-(4.37) and are the average values of the measured currents.
The harmonic analysis of the generator phase voltage and current is shown in Fig.
4.25, when the system has rated power (4000 rpm / 46 kW). The THD of the
voltage is 34.7%, while the THD of the current is 0.6%. Both results are improved
compared to the harmonic analysis of the 3-level drive in Fig. 4.9 that has shown
a phase voltage THD equal to 54.7% and current THD of 1.8% .

87
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

DC-link voltage Vdc tot [V]


1010 4010
Capacitor Voltages [V]

1005 4005
VC1
1000 4000
VC2
995 VC3 3995
VC4
990 3990
1.02 1.04 1.06 1.08 1.1 1.02 1.04 1.06 1.08 1.1
Time [s] Time [s]
(a) (b)

Fig. 4.23: Simulation results of (a) the dc-link capacitor voltages and (b) the total dc-link
voltage when feedforward-based voltage balancing is used

0 10
Real Real
-5 Estimated Estimated
0
Id1b [A]

Id2b [A]

-10
-10
-15

-20 -20
1.02 1.04 1.06 1.08 1.1 1.02 1.04 1.06 1.08 1.1

10 20

15
5
Id3b [A]

Id4b [A]

10
0
Real 5 Real
Estimated Estimated
-5 0
1.02 1.04 1.06 1.08 1.1 1.02 1.04 1.06 1.08 1.1
Time [s] Time [s]

20

15
Id5b [A]

10

5
Real
Estimated
0
1.02 1.04 1.06 1.08 1.1
Time [s]

Fig. 4.24: Simulation results of the dc currents of the grid-side NPC converter at rated
power of the generator (4000 rpm / 46 kW) when feedforward-based voltage
balancing is used

88
4.5. Electric Drive Design using 5-level NPC Converters

20
2000

Generator Current IgenA [A]


Phase Voltage - VgenA [V]
10
1000

0 0

-1000
-10
Vgen a
-2000 Igen a
-20
1.4695 1.47 1.4705 1.471 1.4715 1.472 1.4725 1.473 1.4735 1.474
Time [s]

(a)

2000 5-level generator-side VSC, THDVphGen=34.7%


Phase Voltage VgenA [V]

1500

1000

500

0
0 5 10 15 20 25 30 35 40
Frequency [kHz]

(b)
0.5
5-level generator-side VSC, THD =0.6%
Generator Current IgenA [A]

IGen
0.4

0.3

0.2

0.1

0
0 5 10 15 20 25 30 35 40
Frequency [kHz]

(c)
Fig. 4.25: Simulation results of the 5-level generator voltage and current at rated power
(4000 rpm / 46 kW) when hardware voltage balancing is used: (a) time-domain
analysis and (b)-(c) FFT of the voltage and current

4.5.3 Combined Hardware and SVM-based Voltage


Balancing
The SVM-based voltage balancing method works effectively when the 5-level NPC
converters operate at regions 0 and I of the SVM plane and their modulation
index is lower than 0.5. For higher modulation index, the SVM algorithms needs
to be modified, as shown in Subsection 4.5.1, and the ac output voltage of the

89
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

NPC converters cannot have 5-level shape, since some specific switching states
of the converter are not used. On the other hand, the hardware-based voltage
balancing works effectively under every operating condition, but an additional
dc/dc converter needs to be installed.
One solution would be to combine the advantages of both balancing techniques,
the hardware- and the SVM-based, and build a new balancing converter with
reduced number of components which will operate only when the redundant states
of the NPC converter are not enough to keep the dc-link capacitors under control.
The circuit diagram of the generator drive studied here is shown in Fig. 4.26.
Compared to the diagram in Fig. 4.19, the balancing converter which is used
here has only two stages, regulating the capacitors C1 -C2 and C3 -C4 . The voltage
balancing strategy depends on the modulation index. The grid side converter
is operating at an almost constant modulation index m̂grid ≥ 0.5, whereas, the
generator side converter has variable modulation index m̂gen . The control structure
of the whole electric drive system (an NPC converter connected to the grid and
another converter at the generator side of the power plant) can be summarized in
the flowchart of Fig. 4.27.
When the modulation index m̂gen is at Regions 0 and I, the capacitor voltages

DC-link Balancing Converter


Id 5a Id 5b
Stage 2

VC4 Tb2'
C4
Id 4a Id 4b
L2
Three-phase C3 VC3 Tb2 Grid
Three-phase
PMSG 5-level NPC Id 3a Id 3b Grid
5-level NPC Filter
Generator- N Grid-side
side
C2 VC2 Tb1
Id 2a Id 2b
L1
C1 VC1 Tb1'
Id 1a Id 1b
Stage 1
VC 1÷4
Fig. 4.26: Power circuit of a five-level generator drive with a 2-stage voltage balancing
converter at the dc-link

90
4.5. Electric Drive Design using 5-level NPC Converters

are balanced only from the control system at the generator-side converter and the
grid converter is controlled with conventional SVM [127]. The balancing converter
does not need to operate during that time.
When the generator-side NPC converter is operating at the Region II of the space
vector plane, the redundant states are not enough to control the capacitor voltages
and at the same time to keep 5-level voltage. Therefore, the dc/dc converter at
the dc-link is activated to keep the VC1 equal to VC2 and the VC3 equal to VC4 .
The voltage potential at the middle of the dc-link N is controlled by injecting
zero-sequence voltage as described in the next paragraph.

Region 0  Grid Converter: Standard SVM


0  mˆ gen  0.25 – no balancing
 Generator Converter: Table 4.5

mˆ grid  0.5 Region I  Grid Converter: Standard SVM


0.25  mˆ gen  0.5 – no balancing
mˆ gen  [0, 1]  Generator Converter: Table 4.6

Region II  Grid & Gen. Converter:


0.5  mˆ gen  1 SVM with Zero-Sequence Control
 Hardware Balancing

Fig. 4.27: Flowchart of the dc-link voltage balancing algorithm in a 5-level generator
drive

Zero-sequence injection control (Region II)

The currents of the middle clamping point Id3a and Id3b , are controlled by the
SVM-blocks of both NPC converters using zero-sequence injection. Zero-sequence
injection is used in 3-level NPC converters for balancing their capacitor voltages
in [102] and can also be applied here using a similar control structure. Fig. 4.28
shows the block diagram of the zero-sequence controller that has been implemented
for both space vector modulators of the back-to-back connected NPC converters.
Input of the zero-sequence injection controller is the voltage difference between
the two upper capacitors C3 and C4 and the two lower capacitors C1 and C2 .
This error voltage contains 3rd harmonic components, which are filtered by a
notch filter tuned at three times the fundamental frequency. The filtered signal is
multiplied with the sign of the active power that goes through the NPC converter.
Then, the error voltage goes through a PI controller and the output parameter k
is the input of the SVM.

91
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

m̂ SVM
VC1 with
+ 0
+ 0-seq.
k injection
+ x PI
VC2 + Notch filter for -
limit within
- 3rd harmonic [-1, 1]
VC3
+
PNPC
+
VC4 sign

Fig. 4.28: Block diagram of the zero-sequence controller for regulating the potential at
the middle point N of the dc-link

The switching sequence of the upper triangles from (4.14)-(4.15) for both NPC
converters changes to

(1+k)∙ (1-k)∙ (1-k)∙ (1+k)∙


TE/4 TF/2 TG/2 TE/4 TE/4 TG/2 TF/2 TE/4

E F G E E G F E
Tsw
V'E V'F V'G V'E V'E V'G V'F V'E
+[1 1 1] +[1 1 1] (4.38)

and for the lower triangles

(1+k)∙ (1-k)∙ (1-k)∙ (1+k)∙


TD/4 TE/2 TF/2 TD/4 TD/4 TF/2 TE/2 TD/4

D E F D D F E D
Tsw
V'D V'E V'F V'D V'D V'F V'E V'D
+[1 1 1] +[1 1 1] (4.39)

The combination of software and hardware control techniques presented here


ensures that the additional dc/dc converter is used only when the redundant
states of the SVM are not enough for effective dc capacitor voltage balancing.
Therefore, the efficiency of the whole system is improved compared to the fully
hardware-based balancing of Subsection 4.5.2. Also, the number of components
needed for the dc/dc converter has been reduced, because zero-sequence injection
is used for controlling the middle point current instead of using a 3rd stage of the
balancing converter. Four- and five-level ac voltage is also created when the NPC
converters operate at high modulation index, without the limitations imposed in
Subsection 4.5.1, where the voltage has maximum three levels.

92
4.6. Summary

Simulation results of this voltage balancing technique have been presented in 4

and are not repeated here.

4.6 Summary
This chapter has presented the design process and the simulation analysis of 3-level
and 5-level generator drives. Drives with 5-level converters can have operating
voltage up to 4 times higher than the systems with 2LCs, without using series-
connected power switches. A further advantage of the 5-level converters is the
decrease of the voltage and current harmonics at their ac side, compared to
converters with less number of levels, as shown by the FFT analysis in Figs. 4.9
and 4.25. Simulation results of a 3-level drive have shown that the generator
current and phase voltage THD are 1.8% and 54.7%, respectively. For a drive
system with 5-level converters, the THD of the current decreases to 0.6% and
the THD of the phase voltage to 34.7%, when the same operating conditions are
simulated. The reduction in harmonics that are observed in multilevel drives has
also a positive impact on the performance of the electrical machine, as presented
in 5 where the torque ripple and core losses of a PMSM were drastically reduced
by using multilevel converters.
The voltage drift of the dc-link capacitors is a serious problem in electric drives
with 5-level NPC converters that can be solved by various voltage balancing
strategies. Three effective voltage balancing methods have been described in
Section 4.5 providing also simulation results on a sample 50 kW generator drive for
each of these methods. Each technique has its own advantages and disadvantages,
as described below.
Firstly, it has been shown that the redundant states on the 5-level space vector
plane of the NPC converter’s modulator can be utilized to balance the four dc-link
voltages. This method is an extension of a similar control technique that is also
applied on the 3-level NPC converters, where zero sequence voltage is used to
control the two capacitor voltages of that topology. However, contrarily to the
3-level converter, the redundant states are not enough to control all the four
capacitor voltages when the 5-level drive operates at the outer regions of the space
4 G. Mademlis, Y. Liu, and N. Saadat, "Combined voltage balancing techniques of the DC link

in five-level medium voltage NPC back-to-back converters for offshore renewable generation",
in Proc. 19th European Conference on Power Electronics and Applications (EPE’17 ECCE
Europe), 2017, pp. P.1-P.10.
5 G. Mademlis, Y. Liu, and J. Zhao, “Comparative Study of the Torque Ripple and Iron Losses

of a Permanent Magnet Synchronous Generator Driven by Multilevel Converters,” in Proc. IEEE


XIII International Conference on Electrical Machines (ICEM), 2018, pp. 1406–1412.

93
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

vector plane. One way to overcome this problem, is to switch the converter in
3-level mode avoiding the use of some switching states, as shown in Subsection
4.5.1. The capacitor voltages are kept balanced in that case, but the ac voltage
of the converter has maximum 3-levels, which results to increased harmonics at
the ac side. The main advantage of this balancing technique is that it does not
require any additional hardware for the control of the dc-link voltages, since the
suitable selection of the redundant states of the converter regulates them.
Another way of balancing the dc-link voltages is the use of dedicated dc/dc
converters. This method is the most expensive, since it requires the use of an
additional converter and, thus, the cost, losses and complexity of the system
are increased. The dc/dc converter can either be controlled open-loop by just
switching with constant duty cycle without any feedback of the capacitor voltages
or a more complicated feedforward-based closed-loop control system can be used.
The open-loop control has the simplest structure, while the closed-loop controller
has better dynamic performance and balances the capacitor voltages with smaller
steady state error. The SVM of the NPC converters is not utilized in this case
for the dc-link control, which makes the redundant states of the NPC converters
available to be used in other control strategies, such as in loss and noise reduction
of the ac/dc converter [134].
The third method proposed in this thesis is the combination of the two previous
voltage balancing techniques. The SVM-based voltage balancing method is clearly
advantageous when the converter operates with low modulation index, while it
should be avoided when high modulation index is used. The dc/dc balancing
converter can be activated in this case, while in all the other cases it can remain
idle. Zero-sequence injection can also be used to limit the number of components
of the balancing converter, as described in Subsection 4.5.3. This method is the
most complicated out of the three previously mentioned, since it includes the
control for the additional dc/dc converter as well as the optimal selection of the
redundant states of the modified SVM. However, the 5-level drive can operate
without any limitation, unlike in the first method. Also, the additional hardware
needed is less compared to the second method.
The advantages and disadvantages of the voltage balancing techniques that have
been discussed previously are summarized in the following Table 4.8. The compar-
ison of the system complexity in the table refers to the complexity of the control
in each voltage balancing scheme. The third method has the most advantages,
because it balances successfully the dc-link capacitor voltages, without increasing
considerably the cost of the system, while having higher complexity in the control

94
4.6. Summary

and the hardware design. However, the optimal choice of the voltage balancing
technique depends on the exact needs of the application where the drive is intended
to be used.

Table 4.8: Summary of DC-link Voltage Balancing Techniques

Balancing method Complexity Cost AC Harmonics


SVM-based %
% 0
(Subsection 4.5.1) (3-level voltage)
Hardware-based with open-loop
0 %% 0
control (Subsection 4.5.2)
Hardware-based with closed-loop
% %% 0
control (Subsection 4.5.2)
Combination of SVM and
%% % 0
hardware (Subsection 4.5.3)

95
Chapter 4. Multilevel Converter Topologies and Modulation Techniques

96
Chapter 5

Design and Testing of Medium


Voltage PMSM Drive

5.1 System Overview


The prototype power conversion system developed during this study has two
identical 5-level NPC converters placed at the generator- and the grid-side of the
drive and are responsible for controlling the generator and for the synchronization
to the grid, respectively (Fig. 5.1). There is a third converter connected in parallel
with the dc-link of the whole structure, which is responsible for balancing the
voltages of the dc-link capacitors. The positive flow of energy and the positive
d-, q-currents at the output of the converters are marked with blue arrows on the
schematic. The connections at the ac side of the converters between the PMSG
and the generator-side VSC and between the transformer and the grid-side VSC
are implemented using 3.3-4.2 kV litz-wire cables from Nexans with conductor
cross section 16 mm2 (product code 10148758).
The switches of the power converters are controlled by gate drivers, which send
suitable signals to the gate of the switches, when they need to turn on and off.
The gate drivers receive these control signals from an MCU, as seen in Fig. 5.1.
The MCU and the gate drivers are electrically isolated from each other and this
isolation is implemented by transferring the PWM signals to the gate drivers
through fibre optic cables.
Measurements of the currents and voltages at various points of the power conversion

97
Chapter 5. Design and Testing of Medium Voltage PMSM Drive

3-phase 3-phase
M1 DC-link Grid-side VSC M2
Gen-side VSC
2.8/0.4 kV (+)
Grid Tm (+) M3 Grid
AC DC
5
IM 5

Drive IM PMSG
DC AC
400V YNd11
Igen VC1...VC4 I S’ V S’
θgen ωgen
(as described in DC
Section 2.3) DC
IL
Gate Gate Gate
Drivers Drivers Drivers

MCU and Measurements Interface

MCU
TMS320F28379D

Fig. 5.1: Schematic of the prototype medium voltage power conversion system including
the power converters and the control boards

system are needed in order to control the converters. The generator and grid
currents and voltages are measured with the measurement blocks M1 and M2 . The
voltage of the capacitors at the dc-link is also measured with another measurement
block M3 . The measured signal is sent to the MCU through the "MCU and
Measurements Interface". Three MCU boards from Texas Instruments with
product code TMS320F28379D are used to control the two NPC converters and
the dc-link balancing converter.
The following Section 5.2 describes the design process of the power converters
and the measurement devices and Section 5.3 presents experimental results of the
multilevel generator drive.

5.2 Prototype PMSG Drive Design


The design parameters of the prototype low-scaled power conversion system is
given in Table 5.1. The system described here is a low-power-scaled version, but
with higher operating voltage, of the 500 kW TUSK system shown in Table 3.1.
The prototype presented in this chapter has been used in the thesis as a proof of
concept for the proposed multilevel generator drive and for verifying the developed
control strategies described previously in Chapter 4. The dimensions and the
power density of the converters have not been optimized in order to allow easier
accessibility and measurements with the oscilloscope.
SiC Power switches are used for the converters of the prototype. SiC has ad-
vantageous performance in terms of lower switching losses and higher operating

98
5.2. Prototype PMSG Drive Design

Table 5.1: Parameters of the developed multilevel power conversion system

Parameter Value Unit


Nominal Power Pnom 50 kW
Maximum dc-link voltage Vdctot 4 kV
DC-link capacitance 295 µF
Inductors of the balancing conv. L1 , L2 , L3 4 mH
Maximum ac L2L voltage (rms) 2.8 kV
Phase current (rms) 10.31 A
Switching frequency fsw 20 kHz

temperatures compared to the Si power switches [135, 136]. Thus, SiC switches
allow the construction of more compact converters due to the higher switching
frequency and the resultant decrease of the grid filter size, as well as lower cost for
cooling, because of the lower switching losses and higher allowed temperatures.
The main disadvantage of these switches is that they have at the moment higher
purchase cost compared to the Si IGBTs and MOSFETs. However, their cost is
constantly declining, as the SiC technology becomes more mature.
Suitable application regions for each of the available semiconductor devices are
shown in Fig. 5.2, namely SiC power switches can substitute IGBTs in applications
with power between 1-500 kW and switching frequency higher than 10 kHz. In
applications with even higher power the power switches made of Si are expected to
remain dominant in the future. The graph of Fig. 5.2 was presented by Infineon

Fig. 5.2: Recommended application ranges for Si, SiC and GaN devices [ISiCPEAW 2016
- presented by Infineon Technologies AG 2016]

99
Chapter 5. Design and Testing of Medium Voltage PMSM Drive

(a) (b)

Fig. 5.3: (a) SiC MOSFET SCH2080KE and (b) SiC schottky diode C3D10170H

Technologies AG at the ISiCPEAW workshop in 2016. A proof-of-concept 50 kW


inverter of Wolfspeed [137] shows that SiC power switches can reduce the total
system cost by 15% compared to Si power switches, if the cost of cooling, filters,
etc. is considered. More specifically, although the cost of the power devices is
much higher compared to a Si-based system, the cost and size of the grid filter
inductors and of the cooling system can be reduced in a SiC converter. This is
particularly useful in the case of the undersea kite, since a size reduction of the
passive filter at the ac side of the power conversion system is highly beneficial.
The SiC MOSFETs SCH2080KE of ROHM [Fig. 5.3(a)] are used for the lab
prototype of the converters, which have been modelled in the simulation results
of Chapter 4. These switches have maximum drain-source voltage 1200 V and
maximum continuous drain current 40 A (at 25°C) / 28 A (at 100°C). The clamping
diodes of the NPC inverters are the C3D10170H from CREE [Fig. 5.3(b)] with
1700V maximum reverse voltage and 29 A (at 25°C) / 14.4 A (at 135°C) maximum
continuous current.
Voltage overshoots caused by parasitic capacitance and inductance between the
switching devices are a serious problem, when designing a power converter. There-
fore, the power switches should not operate very close to their absolute rated
voltage, but a safety limit of approximately 20% or more should be kept. The
exact number of the safety margin depends eventually on the design of the power
conversion system. The nominal dc-link voltage of the power conversion system is
designed to be equal to 4 kV and, therefore, each switch of the converters will have
to withstand 1 kV. In this case, a safety margin of 200 V has been chosen between
the nominal operating conditions and the maximum voltage of the switches.

100
5.2. Prototype PMSG Drive Design

5.2.1 SiC Gate Driver Design


The gate driver is responsible for transferring the control signal from the output
of the ePWM peripheral of the MCU to the gate of the MOSFETs. They also
provide electrical isolation between the high voltage potential of the switch and the
low voltage potential of the microcontroller. The SCH2080KE MOSFET requires
a gate-source voltage of 20 V in order for the switch to turn on and -2.7 V for the
switch to turn off.
Fig. 5.4(b) shows the circuit diagram of the gate driver PCB. Two separate dc/dc
converters are used for each gate driver board which provide ±15 V and ±5 V
power supplies. The primary side of these converters receives 24 V input voltage.
The secondary side of the dc/dc converters is floating at the voltage potential of

(a)

Drain
HFBR-2522Z Gate Driver
3.3V IXDD609PI SCH2080KE
Iout = 9A (peak)
-2.7V D3 20Ω Gate

20Ω Distance
20V -2.7V
DC
LED1
24V 5V D1
Vin 24V 2.7V
DC -10V
1.1kΩ 3.3V 470Ω 47kΩ
15V
D2
Linear Regulator 20V
750Ω
10 kV isolation (Positive)

5V
DC 278.4Ω -2.7V Source
24V Ground
DC 240Ω
-5V
5V -5V

Linear Regulator
10 kV isolation (Negative)

(b)

Fig. 5.4: (a) Picture and (b) schematic of the gate driver for the SCH2080 SiC MOSFET

101
Chapter 5. Design and Testing of Medium Voltage PMSM Drive

the source of the MOSFET, while the primary side is at the voltage potential of
the MCU. Therefore, dc/dc converters with reinforced isolation voltage of 10kV
need to be used.
Linear regulators modify the output voltage of the dc/dc converters in order to
produce −2.7 V and 3.3 V in respect to the ground, which is the source of the
MOSFET. The gate resistor is equal to 20 Ω for turning on the switch and 10 Ω
for turning off. Lower resistance is used for the turn-off operation in order to
increase the speed of the gate driver. This is implemented by using a schottky
antiparallel diode D3 in parallel to the turn-on resistor [138]. Zener diodes D1 and
D2 and a large resistor of 47 kΩ are used between the gate and source outputs of
the gate driver board in order to protect it from overvoltages.
The distance between the outputs of the gate driver IC up to the gate and source
of the switch is very critical for the optimal design of both the gate driver and the
inverter PCB. Parasitic inductance occurs due to the loop between the gate driver
output and the ground return traces as can be seen on the right side of Fig. 5.4(b),
which can create voltage ringing in the gate drive waveform [138]. The voltage
ringing is defined as unwanted oscillations during a step change of a voltage signal,
because the parasitic inductance does not allow the current flowing on that loop
of the system to change rapidly its value. Ringing on the power switch can cause
unwanted turn-on and turn-off of the switch, which creates additional switching
losses and could eventually destroy the switch if these voltage oscillations surpass
its maximum voltage ratings. The longer this distance is, the larger the parasitic
inductance becomes and more severe oscillations can occur on the power switch.
The voltage oscillations due to the ringing usually have high frequency at MHz
level, which create EMI noise that can easily interfere with other sensitive devices
of the system, such as measurement devices and the MCU.
In order to decrease these oscillations, the following solutions are available to the
designer of the PCBs. First, the distance between the gate driver and the power
switch has to be minimized. A wider trace for the gate signal can also limit the
parasitic inductance. Surface mounted devices (SMD) are also preferable compared
to through-hole devices, since they can deal better with EMI and their reduced
physical dimensions can limit the parasitic inductance and allow closer placement
of the components [139]. In case the problem cannot be solved, capacitors have
to be placed close to the gate pin of the switch and on the gate driver PCB in
order to filter out the unwanted oscillations. The filter frequency can be optimized
by placing multiple capacitors of different capacitance value, which have different
resonant frequency.

102
5.2. Prototype PMSG Drive Design

Fig. 5.5(a) shows the experimental measurements of the gate to source voltage VGS
at the terminals of the switch with switching frequency of 5 kHz. The switching
frequency of the prototype converters used in the experimental results in later
sections is equal to 20 kHz. A closer look of the VGS voltage can be seen in Figs.
5.5(b)-(c), which validate that the voltage created by the gate driver is within the
recommended range mentioned in the MOSFET datasheet [140], namely maximum
VGSS−surge ∈ [−10, 26] V.

30

20
VGS [V]

10

-10
-0.2 -0.1 0 0.1 0.2
Time [ms]
(a)

25 25 X: -0.09947
X: -0.1994 Y: 19.5
20 Y: 23.02 20
X: -0.1989
15 Y: 19.82 15
VGS [V]
VGS [V]

10 10
5 5
X: -0.2002 0 X: -0.09746
0 Y: -3.54
Y: -3.54
-5 -5 X: -0.09921
-0.2005 -0.2 -0.1995 -0.199 -0.101 -0.1
Y: -4.5
-0.099 -0.098 -0.097
Time [ms] Time [ms]
(b) (c)

Fig. 5.5: Experimental results of (a) the gate-source voltage of SCH2080KE MOSFET
with switching frequency of 5 kHz; zoomed-in gate-source voltage (b) during
turn-on and (c) turn-off transient

5.2.2 NPC Converter Board


The developed prototype of the NPC inverter is shown in Fig. 5.7. The inverter
has nominal voltage higher than 1.5 kV peak voltage and, therefore, some special
rules are followed during the design process. The prototype PCBs have been built
based on the design guidelines described in [141].
The first thing that needs to be considered is the proper substrate material for
the PCB, especially if double-layer PCBs are used (Fig. 5.6). PCBs with more
than two layers are not recommended for high power converters, since they can
become fragile due to the large thickness and it is also difficult to dissipate the

103
Chapter 5. Design and Testing of Medium Voltage PMSM Drive

heat from the inner copper layers [142]. The substrate of the PCB needs to be able
to handle the differential voltage between the top and the bottom copper layer.
The substrate (prepreg) used on the developed prototype by the manufacturer of
the PCBs is the FR-370HR with dielectric strength of 54 kV/mm [143].
Proper design of the traces on the PCB is very important in order to avoid Corona
discharges or a direct arc. Sharp edges and corners on the copper traces should
be avoided, because high electric field occurs there, which can initiate Corona.
Smooth and round curves should be used wherever possible for the traces and
the pads of the PCB [141]. Also, the minimum creepage distances should be
kept between the components placed on the PCB and between the traces of the
same layer. A rule of thumb of 100 V/mm for the creepage distance between the
components has been considered for the design of the prototype.
Due to the high operating voltage of the prototype, longer distances are, therefore,
required between the power switches of the converters and the dc-link capacitors
of the system. Since the dc-link capacitors operate as a quite stiff voltage source,
the parasitic inductance of the PCB traces can cause high frequency voltage
oscillations of the switches, which is called ringing. Proper design of the traces in
the PCB is very important, as mentioned also in Subsection 5.2.1. The parasitic
inductance of the circuit can be reduced in order to limit the ringing by increasing
the width of the traces of the PCB as much as possible. Also, the drain and
source traces of the power MOSFETs as well as the anode and cathode traces
of the schottky diodes should be placed directly over each other, which reduces
the parasitic impedance [139]. Resistor-capacitor (RC) snubber circuits are also
installed in parallel to each MOSFET and diode of the NPC converter, as can be
seen in Fig. 5.7(a). The RC snubber consists of a small capacitor and a small
resistor connected in series that suppress the rapid rise in voltage (dV/dt) of the
switching device to a value that will not trigger ringing.

Silkscreen
Soldermask
Copper
Substrate

Fig. 5.6: Composition of a two-layer PCB

104
5.2. Prototype PMSG Drive Design

(a)

(b)

Fig. 5.7: PCB of a single phase 5-level NPC converter with the gate drives connected to
the switch terminals: (a) top and (b) bottom view

5.2.3 DC-Link Board


Film capacitors with part number 947D591K132DJRSN are chosen for the proto-
type power conversion system, which have rated voltage of 1.3 kV and capacitance
of 590 µF. Film capacitors are more bulky and expensive compared to electrolytic
ones. However, they can handle higher voltage and are, therefore, more commonly
used in high voltage drives. Eight capacitors (in two parallel-connected groups of

105
Chapter 5. Design and Testing of Medium Voltage PMSM Drive

Fig. 5.8: Laboratory prototype of the three-phase 5-level NPC inverter

C1 C2

Connections to
other dc-link
Voltage boards
measurements
board

Phase A C4 Phase B C3 Phase C

Fig. 5.9: Schematic of the dc-link PCB design, red→top copper layer and blue→bottom
copper layer

4 capacitors) are installed in total on the PCBs of the power converters through a
special board which is shown in Figs. 5.8-5.9. Therefore, the equivalent capacitance
of the dc-link becomes 590 · 24 = 295 µF. The rated current of each capacitor is
79 A(rms), which is enough for the specific project.
This dc-link PCB serves as the base for connecting the three-phase NPC inverter
and the balancing converter to the dc-link of the system. As can be seen in Fig.

106
5.2. Prototype PMSG Drive Design

Phase c DC-link Phase c


Generator Balancing Converter Grid
Phase b PCB Phase b
Converter Phase a Phase a Converter
TM1
Tb3' TG1
C4 VC4
TM2
D1 D1 TG2
TM3
D2 L3 D2 TG3
C3 VC3 Tb2 Tb3
TM4
D3 N D3 TG4
PMSG Grid Grid
TM5 D4 L2 D4 TG5
C2 VC2 Tb2' Tb1'
TM6
D5 D5 TG6
TM7 L1
D6 C1 VC1 D6 TG7
Tb1
TM8 TG8

Fig. 5.10: Schematic diagram of the whole prototype power conversion system

5.9, this board has enough connectors for 3 converter boards and 4 capacitors.
Therefore, three of these PCB are needed to accommodate the two B2B-connected
NPC converters and the balancing dc/dc converter [Fig. 5.18(b)]. Fig. 5.10 shows
also the schematic diagram of the whole power conversion system, which is the
same with the schematic of the simulated system in Fig. 4.19.
It is important to make sure that the distance between the dc-link terminals of the
power converters and the pins of the capacitors is the shortest possible and that
the traces on the dc-link PCB are wide in order to keep the parasitic inductance
low. The traces connected to the plus and negative poles of each capacitor are also
placed over each other, as explained also in Subsection 5.2.2, which can reduce
the parasitic inductance even further.

5.2.4 Measurement Devices


Voltage measurements of the four dc-link voltages and of the three phase voltages
at the grid output of the converter are needed for the control of the system. The
Hall-effect voltage sensors from LEM with product number LV 25-P are used,
as can be seen in Fig. 5.11. The voltage to be measured is shown as V1 in Fig.
5.11(b). The nominal primary current of the sensor is 10mA; therefore, a suitable
resistor R1 is placed between the node V1 and the input of the sensor HV+. All
the required voltages are measured in respect to the neutral point of the dc-link N
and, therefore, the negative pole of the sensor HV- is connected to N . The neutral

107
Chapter 5. Design and Testing of Medium Voltage PMSM Drive

(a)

LV 25-P
IPN : ISN = 10 : 25
V1 R1
± 10 mA
HV+
VM
ADC
VM΄
VCC [0, 3] V [0, 4095]
N 12bit
HV- +15V

RG RM 1.5V
M ± 25 mA for ac
measurements
for dc
-15V measurements
Vshift

(b)

Fig. 5.11: (a) Laboratory prototype and (b) schematic diagram of the dc and ac voltage
measurement PCB

point N is grounded though a ground resistor RG = 14.1 kΩ.


The LV 25-P sensor behaves as a current source at its secondary side with nominal
secondary current 25 mA. The voltage VM ∈ [0, 3] V across the resistor RM is the
input of the ADC channel of the MCU. When the voltage to be measured V1 is ac
and gets both positive and negative values, the signal VM needs to be shifted up
by 1.5 V and the reference of the signal is the 1.5 V, as shown on the right side of
Fig. 5.11(b). In case a dc voltage is being measured, the reference of VM is the
ground. The RM is equal to 60 Ω, when the measured voltage is ac, and is equal
to 120 Ω, when measuring dc voltages.
The ADC of the MCU converts the VM into a 12bit digital signal and the original

108
5.2. Prototype PMSG Drive Design

voltage V1 can be calculated as follows


 0 
3VM
V1 = − Vshif t · gain + Vof f set (5.1)
4095

where the variable gain depends on the ratio between the primary and secondary
side of the measurement device and on the resistor RM .
Similar Hall-effect sensors are used for measuring the currents of the system. More
specifically, the following current measurements are needed (see Fig. 5.1): three-
phase currents of the PMSG Igen , three-phase grid currents Is and the current of
the three dc inductors IL . The current transducer LA 25-P from LEM is used for
the current measurements and similar methodology to the voltage sensors is used
for receiving the measured signal in the MCU.

5.2.5 Overcurrent-Overvoltage Protection Strategy


The signals from all the measurement devices end up on the PCB shown in
Fig. 5.12 before being sent to the ADC inputs of the MCU. This PCB converts
all the measured signals into the correct voltage level which is required for the
ADC inputs. Another use of this board is to disconnect the converter in case an
overcurrent/overvoltage is measured in the system. The analog circuit shown in
Fig. 5.13 compares each measured signal VM with a maximum and minimum
value Vref H and Vref L , respectively. In case VM exceeds the predefined operating
region, an error signal is created, which turns on also a red LED. The global error
signal PRT is the output of this board.
This error signal forces all the power switches of the converters to open, in case
of a fault. This is achieved through the trip-zone module of the MCU [144] and
also though another PCB marked with red color in Fig. 5.13, whose function is to
convert the ePWM outputs of the microcontroller into optical signals, which are
sent to the gate drivers. The PRT disables the PWM signals of the gate drivers
protecting the converters. The error signal is also connected to a relay coupled to
a mechanical switch which switches-off the converters. The detailed schematics of
the PCBs shown in Fig. 5.13 can be found in Appendix B.
Therefore, there is a three-layer overcurrent/overvoltage protection strategy of the
generator drive. The analog circuit that disconnects the control signals of the gate
drivers has the shortest reaction time, since the logical gates have delay in the
range of a few µs.

109
Chapter 5. Design and Testing of Medium Voltage PMSM Drive

Fig. 5.12: PCB collecting all the measurement signals of the drive before routing them
to the MCU

Hardware Protection
Reference signal generator
ePWM
4 MHz trip zone
L12
VRef H1

VRef H2

VRef H3
VRef L1

VRef L2

MCU
VRef

+C
1
VM1 VRef H1 - JK flip-flop
SET

VM2 OR1 J Q
VRef PRT = 3V: No error
VM3 L1 +C
2 PRT = 0V: Error
-
3.3V K
CLR
Q D1
PWM PCB
VM12 PRT HFBR-1522Z
NOR
+C24
C ANODE
NAND1
VRef H12 -
23
PWM1 CATHODE
SET D N.C.
OR12 J Q error N.C.

VRef +C
L12
24 signals PWM2
NAND2
-
3.3V K
CLR
Q D12
LM339D
SN74HC109D NAND3
PWM3

(from MCU)

Fig. 5.13: Schematic diagram of the overcurrent/overvoltage protection strategy

110
5.3. Experimental Results

5.3 Experimental Results

5.3.1 Medium Voltage PMSG


A medium voltage PMSG has been used within this project and it is tested in
this subsection in order to evaluate its performance and measure its parameters.
The electrical machine test bench is shown in Fig. 5.14(e), where the PMSG and
the driving motor IM are at the right and left side, respectively. Pictures of the
generator during its manufacturing phase are shown in Fig. 5.14(a)-(d).
The design of the generator is not part of this thesis.

No-load test of the PMSG

The PMSG is tested first at no-load conditions in order to evaluate the back-
EMF and the cogging torque and, afterwards, load-tests are implemented. The

a b

d
4
1
2

e 3

Fig. 5.14: PMSG (a)+(b)Stator windings, (c) stator, (d) rotor and (e) the whole machine
test-bench: 1) IM, 2) the medium voltage PMSG mounted on the bench, 3)
connection box for the stator cables of the PMSG, 4) terminal for connecting
the encoder and temperature sensors of the PMSG

111
Chapter 5. Design and Testing of Medium Voltage PMSM Drive

following voltage probes are used for the following experiments of this section: the
HVD3106A voltage probe with up to Vpeak =1500 V measurement range [7] and
the CP150 current probe [8].
No-load test of the PMSG is implemented by having the terminals of the generator
open-circuited and operating the IM in motor mode. The measured L2L voltage is
equal to the no-load back-EMF of the machine, which is shown in Fig. 5.15(a) for
rotor speeds equal to 1600 rpm up to 1200 rpm. Fig. 5.15(b) shows one mechanical
period of the no-load EMF voltage and the cogging torque for rotor speed 200 rpm.
The torque waveform has 12 positive and negative peaks in one mechanical period
of the PSMG, which is the 12th harmonic generated by the slots and poles of the
PMSG. There is also 6th harmonic component in the torque waveform generated
by the cogging torque of the IM.

2000 200 300


ngen = 1600 [rpm] VBC [V]

Torque 10-3 [Nm]


1000 ngen = 1400 [rpm] Tm [m Nm] 200
100
Voltage [V]

ngen = 1200 [rpm] 100


VBC [V]

0 0 0
-100
-1000 -100X: 0.04716
Y: -154 -200
X: 0.09854
-2000 -200 Y: -153.2 -300
0 0.005 0.01 0.015 0.02 0 0.05 0.1 0.15 0.2 0.25 0.3
Time [s] Time [s]
(a) (b)

Fig. 5.15: Experimental results of the medium voltage PMSG no-load test: (a) no-load
EMF voltage for various rotor speeds and (b) cogging torque for 200 rpm
rotational speed

Load test of the PMSG without converter control

The PMSG is tested also with a resistive load RL , as shown in Fig. 5.16. The load
is three resistor banks connected in Y connection and can vary its resistance from
0.4 Ω up to 3 Ω. The PMSG is connected to the resistive load through a step-down
transformer and its active and reactive power in each case are determined by the
combination of the equivalent resistance and inductance of the load, since it is not
controlled by a converter.
Experimental results of this test are presented in Fig. 5.17 for two different
operating points. In Figs. 5.17(a)-(b) the load is RL = 3 Ω with ngen = 2000 rpm
and the apparent power is S = 6.42 kVA, while in Figs. 5.17(c)-(d) the load is

112
5.3. Experimental Results

2.8/0.4 kV RL
Grid Tm (3-ph. Y)
IM
Drive IM PMSG
400V
YNd11
Igen Vgen
θgen ωgen
(as described in
Section 2.3)

Fig. 5.16: Schematic diagram of the PMSG load test with direct connection of the
generator to the resistive-inductive load

RL = 0.8 Ω with ngen = 2700 rpm and the apparent power is S = 12.27 kVA.
The parameters of the PMSG are shown in Table 5.2, where the inductances in
the d- and q-axis have been measured based on the method described in [145, 146].
Specifically, step-change of dc current was applied towards the d- and q-axis of
the machine and the inductance was calculated from the response of the current.

15
10 1000
5
Vgen [V]
Igen [A]

0 0
-5
Ia Vbc
-10 -1000
Ib Vca
-15
0 0.005 0.01 0.015 0.02 0 0.005 0.01 0.015 0.02
Time [s] Time [s]
(a) RL = 3 Ω, ngen = 2000 rpm (b) RL = 3 Ω, ngen = 2000 rpm
15
10 1000
5
Vgen [V]
Igen [A]

0 0
-5
Ia Vbc
-10 -1000
Ib Vca
-15
0 0.005 0.01 0.015 0.02 0 0.005 0.01 0.015 0.02
Time [s] Time [s]
(c) RL = 0.8 Ω with ngen = 2700 rpm (d) RL = 0.8 Ω with ngen = 2700 rpm

Fig. 5.17: Experimental results of the medium voltage PMSG with resistive-inductive
load

113
Chapter 5. Design and Testing of Medium Voltage PMSM Drive

Table 5.2: Rated values and control parameters of the PMSG and the IM

Parameter Value Unit


Nominal Power Pnom 50 kW
Nominal rotor speed 4000 rpm
Number of pole pairs p 6 -
Nominal torque TN 119 Nm
No-load EMF at rated speed (L2L, peak) 2932 V
Stator inductance Ld 18.9 mH
Stator inductance Lq 25 mH
Stator resistance RS 1.5 Ω
PMSG inertia JP M SG 0.1274367 kg·m2
IM inertia JIM 0.2 kg·m2
Torque sensor inertia JS 175.8·10−6 kg·m2
PMSG current controller bandwidth αc 1500 rad/s
PMSG speed controller bandwidth αω 15 rad/s
PMSG converter switching frequency fsw 20 kHz

5.3.2 Power Conversion System


The multilevel generator drive is tested in this subsection and the laboratory
set-up is shown in Fig. 5.18(a) with the electrical machine test-bench and the
power converter cabinet. Fig. 5.18(b) shows a more detailed view of the power
converter cabinet. The converters installed inside the cabinet are marked with
dashed boxes and are described in the figure caption.
The dc-link voltage balancing strategy of Subsection 4.5.2 is used here, namely
the three-stage balancing converter with open-loop control of its duty cycle. The
balancing converter has the same switching frequency with the NPC converters,
namely 20 kHz.

Generator-side Converter

Experimental results of the generator-side NPC converter are shown in Figs. 5.19-
5.20. The PMSG operates in generator-mode with 650 rpm and has reference
current equal to Igenq

= −4 A. The total dc-link voltage is 670 V during this test.
The FFT analysis of the voltages and currents is also shown and specifically the
THD of the current is 3.997%, while the THD of the phase voltage is 37.338%.
When compared to the simulation results of Fig. 4.25 in Subsection 4.5.2, it can

114
5.3. Experimental Results

(a) (b)

Fig. 5.18: (a) Laboratory set-up of 1) the MCU and measurements board, 2) the power
converter cabinet, 3) the 2.8/0.4 kV three-phase transformer 4) the machines
test-bench; (b) closer view of the power converter cabinet with 1) the grid
side 5-level converter, 2) the dc-link balancing converter, 3) the generator-side
5-level converter and 4) the dc inductors for the balancing converter

5
Generator Voltage [V]

Generator current [A]

X: 403.7 X: 419
500 Y: 698.2 Y: 697.2

0 0
VgenAB
IgenA
VgenBC IgenB
-500 IgenC
VgenCA
-5
400 410 420 430 440 450 400 410 420 430 440 450
Time [ms] Time [ms]

(a) (b)
Fig. 5.19: Experimental results of the generator-side NPC converter when the generator
operates at 650 rpm and current Igenq

= −4 A: (a) three-phase L2L generator
voltage and (b) current

be seen that the phase voltage THD is very close to the simulated value. However,
the THD of the current is higher at the experimental results and this is explained
due to the lower amplitude of the current there.

115
Chapter 5. Design and Testing of Medium Voltage PMSM Drive

100
THDVgenAB = 16.395%
Vgen AB [V] 80

60

40

20

0
0 5 10 15 20 25 30 35 40
Frequency [kHz]

(a)
100
Phase Voltage Vgen AN [V]

THD = 37.338%
VgenAN
80

60

40

20

0
0 5 10 15 20 25 30 35 40
Frequency [kHz]
(b)
0.15
THDIgenA = 3.997%

0.1
Igen A [A]

0.05

0
0 5 10 15 20 25 30 35 40
Frequency [kHz]
(c)
Fig. 5.20: Experimental results of the generator-side NPC converter when the generator
operates with 650 rpm and current Igenq

= −4 A: FFT analysis of the (a) L2L
voltage VgenAB , (b) phase voltage VgenAN and (c) stator current IgenA

Grid-side Converter

Experimental results of the grid-side converter are shown in the following figures.
The three-phase transformer shown in Fig. 5.1 is not used here, because low
voltage tests are implemented. Instead, the grid converter is connected to the
400 V-grid through a series-connected inductor with inductance Lf = 2 mH. The
grid-side converter control parameters are shown in Table 5.3.
The current controller of the grid current Is is tested and experimental results

116
5.3. Experimental Results

Table 5.3: Parameters of the Grid-side Converter

Parameter Value Unit


Tested L2L grid voltage (rms) Vs 400 V
Fundamental frequency fn 50 Hz
Grid converter switching frequency fsw 20 kHz
Grid filter inductance Lf 2 mH
Tested total dc-link voltage Vdc tot 670 V
Grid current controller bandwidth αcs 8000 rad/s
PLL bandwidth αP LL 10π rad/s
DC-link controller bandwidth αDClink 400 rad/s

6 500 6
Grid Voltage VsConv [V]
Grid Current Is [A]

Grid Current Is [A]


4
4
2
2 0 0
-2
0 I*sd VsConvAN
Isd IsA -4
-2 -500 -6
90 95 100 105 110 50 100 150
Time [ms] Time [ms]
(a) (b)
2 500 6
Grid Voltage VsConv [V]
Grid Current Is [A]

Grid Current Is [A]


4
0
2
-2 0 0
-2
-4 I*sd VsConvAN
Isd IsA -4
-6 -500 -6
90 95 100 105 110 50 100 150
Time [ms] Time [ms]
(c) (d)
Fig. 5.21: Experimental results of the grid-side NPC converter with a step-change of the

Isd reference current (a)-(b) from 0 to 5 A and (c)-(d) from 0 to -5 A

of the phase A current IsA and the phase voltage at the output of the converter
VsConvAN are shown in Figs. 5.21-5.22. The grid phase voltage is defined as
the differential voltage between a phase of the converter and the middle point of
the dc-link, marked as N . Specifically, step-change of the Isd

is applied and the
converter is sending active power to the grid in Fig. 5.21(a)-(b), while it receives
power from the grid in Fig. 5.21(c)-(d). Similarly, q-axis current Isq

is applied in
Fig. 5.22 and the grid converter is exchanging reactive power with the grid.

117
Chapter 5. Design and Testing of Medium Voltage PMSM Drive

6 500 6

Grid Voltage VsConv [V]


Grid Current Is [A]

Grid Current Is [A]


4
4
2
2 0 0
-2
0 I*sq VsConvAN
Isq IsA -4
-2 -500 -6
90 95 100 105 110 50 100 150
Time [ms] Time [ms]
(a) (b)
2 500 6

Grid Voltage VsConv [V]


Grid Current Is [A]

Grid Current Is [A]


4
0
2
-2 0 0
-2
-4 I*sq VsConvAN
Isq -4
IsA
-6 -500 -6
90 95 100 105 110 50 100 150
Time [ms] Time [ms]
(c) (d)
Fig. 5.22: Experimental results of the grid-side NPC converter with a step-change of the

Isq reference current (a)-(b) from 0 to 5 A and (c)-(d) from 0 to -5 A

The dc-link controller of the grid-side NPC converter is tested in Fig. 5.23 with a
step-change of the reference voltage, which changes from 670 V to 600 V at the
time instant 85ms. The individual voltages of the four dc-link capacitors in Fig.
5.23(b) follow the change of the total dc voltage. They remain balanced with each
other with a small error during this transient. This error can be explained due to
the use of open-loop control for the balancing converter, as mentioned previously.
Therefore, larger errors between the capacitor voltages are expected compared to
the closed-loop control scheme shown in the simulation results of Subsection 4.5.2.
The amplitude of the L2L grid voltage at the output of the converter VsConvAB
changes as well, as can be seen in Fig. 5.23(c), since it varies between +Vdc tot
and −Vdc tot . The converter voltage is approximately 40 V higher than the dc-link
voltage due to the ringing on the power switches. The grid voltage VsAB is also
shown in the same figure, which is in phase with the converter voltage. The
d-axis grid current Isd in Fig. 5.23(d) is calculated by the dc-link PI controller.
A limiter of 6 A is applied on the reference d-current, which is reflected on the
actual current.

118
5.4. Summary

DC-link Voltage Vdc tot [V]

Capacitor Voltage VC [V]


700 X: 84.85
170
Y: 671.6 X: 84.95 VC4
Y: 168.2
VC3
650 160 VC2
X: 148.8 VC1
Y: 596.1
600 150
X: 148.5
Y: 148.9

550 140
50 100 150 50 100 150
Time [ms] Time [ms]
(a) (b)
10

Grid Current Isd [A]


X: 65.23 X: 85.87
X: 145.6
Grid Voltage [V]

500 Y: 713.6
Y: 635.4
Y: 6.29

5
0
0 X: 85.38
VsConvAB Y: 0.5703
-500
VsAB
-5
50 100 150 50 100 150
Time [ms] Time [ms]
(c) (d)
Fig. 5.23: Experimental results of the grid-side NPC converter with a step-change of the
dc-link voltage Vdc tot from 670 V to 600 V: (a) dc-link voltage, (b) dc-link
capacitor voltages, (c) grid voltage and (d) d-axis grid current

5.4 Summary
In this chapter, the design process and test results of a novel 5-level NPC-based
PMSG drive have been presented. Firstly, the hardware design and component
selection of the NPC converters and the balancing circuit are described. Special
care has been taken so that the power converters are designed with the potential
to operate at medium voltage level, with up to 4 kV at the dc-side. Therefore,
large creepage distance has been kept between the power components in order to
avoid failures due to the high operating voltage and the control and measurement
boards are highly insulated from the power circuit of the converters. The parasitic
inductance on the converter boards is also minimized as much as possible and
snubber circuits are used in parallel to some of the power switches to further reduce
their voltage overshoot during the switching transient. The prototype multilevel
drive uses 1.2 kV SiC power switches in order to decrease the converter losses.
Suitable gate drive boards have been designed for the specific switches.
Experimental results of both converters of the drive are shown in order to demon-
strate the functionality of the system under various operating conditions. DC-link
voltage up to 670 V and current of 5 A have been tested in this chapter.

119
Chapter 5. Design and Testing of Medium Voltage PMSM Drive

120
Chapter 6

Power-HIL Testing of High-Power


Electrical Machine Drives

6.1 Electrical Machine Emulator


Testing VSCs for electrical machine drives on a machine test-bench, as has been
shown previously in Chapter 5, entails difficulties in case the stability of the
VSC controller has not been verified yet and when operation close to the voltage,
current and speed limit of the system needs to be tested. Safety concerns and
cost limitations for developing bulky electrical machine test benches are further
common issues. The electrical machine might also not be available at the time
when the test has been scheduled.
The electrical machine under-test can be replaced with another VSC that acts as
a virtual load emulating the electrical machine operation. An inductive device,
such as a three-phase inductor or a transformer, needs also to be placed at the
ac coupling between the two VSCs in order to filter the current harmonics. This
testing procedure is called Power Hardware-in-the-loop (P-HIL) and its main
advantage is that many different machine types of similar rated parameters can be
tested using mostly the same hardware, with minor modifications in the control of
the two VSCs. Therefore, experimental testing of the inverter-under-test (IUT)
can be accelerated, reducing at the same time their development cost.
There are two basic configurations of electrical machine emulators. In the first
configuration the IUT and the machine-emulator VSC have separate dc power
supplies with power rating equal to the maximum power that needs to be tested,

121
Chapter 6. Power-HIL Testing of High-Power Electrical Machine Drives

as shown in [147–149]. The main disadvantage of this concept is the requirement


for two high-power supplies, which entails higher cost and complexity. In the
second configuration, dc coupling of the emulator VSC and the IUT can be used
as in [150–152], so that the energy of the system circulates in this loop and only
the power losses of the converters and the passive components need to be supplied
by an external dc power supply.
The second type of machine emulators has the simplest system configuration, due
to its reduced number of components. However, when a three-phase inductor is
used as a filter at the ac coupling of the topology, a conductive path with very low
impedance for the zero-sequence is created. Since the dc link of the two converters
is directly connected in this specific machine emulator, a circulating zero-sequence
current can be created that may reach high values. This circulating current acts
as a disturbance for the current controller of the IUT and introduces additional
load for the power switches. Its value increases, proportionally with the dc voltage
of the topology and the voltage difference between the two inverters. The zero-
sequence current contains low- and high-frequency components [152, 153], with the
low-frequency ones being mainly dc- and third-harmonic and the high-frequency
components being at multiplies of the converters’ switching frequency.
Control techniques for suppressing the circulating zero-sequence current are pre-
sented in this chapter and verified experimentally on a PMSM emulator.

6.1.1 Machine Modelling Approaches


Emulation accuracy of the studied electrical machine depends to a large extend
on the accuracy of the machine model itself, which is implemented either with
analytical equations or/and utilizing simulation data in order for the emulator
system to perform similarly with the physical machine drive.
Simplifications in the modelling process of the studied electrical machines are
often observed in the literature during the development of machine emulators. A
synchronous motor has been emulated in [150], where non-salient rotor structure
is assumed and magnet saturation has not been considered. Similar assumptions
have been made for an induction motor emulator in [154]. This can lead to large
deviations, especially at high-load conditions where the machine core is usually
saturated. This issue can be solved by utilizing physics-based machine model
[148, 155], where the PMSM parameters are determined at any time instant as a
function of the instantaneous current and rotor position through look-up-tables
with FEM data. Look-up-table based machine models are used also in [147],

122
6.1. Electrical Machine Emulator

where the measured dq-axis voltage of the IUT is the input of the model and the
estimated machine current is the output. High-bandwidth current controllers and
use of FPGA is necessary to implement this machine emulator.
Another aspect of study is the type of control used for the machine emulator VSC
in order to generate the correct PWM voltage that would emulate the physical
machine, which are briefly shown in Fig. 6.1. The standard method is to use
closed-loop current control for the emulator VSC. The measured three-phase
voltage of the IUT ve abc is the input of the controller and the reference voltage
for the emulator VSC vem abc is the output variable. However, this would cause a
conflict with the current controller of the IUT, since both inverters have the same
current flowing through their ac terminals. This issue can been solved by using
PI-regulators with higher bandwidth for the emulator-side current controller that
would respond faster than the IUT-side controller [147]. In order to ensure faster
dynamics for the emulator converter, significantly higher switching frequency is
required, which can be three or more times the switching frequency of the IUT [156].
Increasing the switching frequency so much can be challenging for high-power
industrial inverters, which can be alleviated by applying sequential switching [156]
or using SiC devices that have lower switching losses [149]. High-bandwidth model
predictive controller can also be used for current control of the emulator converter
[149]. FPGA is also necessary in these cases to implement the control and model
the machine current with high fidelity emulating even the current harmonics of
the IUT.
The "inverted" machine model is another control approach for the machine em-
ulator, where the emulator converter is controlled in open-loop mode. To avoid
the need for a high-bandwidth current controller, the reference voltage for the
emulator converter that represents the voltage response of the machine is calculated

3-phase Ze 3-phase
3-phase Ze 3-phase
machine machine
IUT Electrical
emulator ve abc IUT
emulator
*
model:
ve abc v* ie abc v (2.7)-(2-8)
ve* abc em abc ve* abc
em abc

Current Machine Torque Mechanical


Speed Current Speed Current
controller controller Controller model calculation: model: (2.4)
ie abc (high-bandwidth) (FEM) controller controller (2.5)

(a) (b)

Fig. 6.1: Simplified control diagram of the machine emulator set-up using (a) individual
high-bandwidth current control for the emulator VSC or (b) open-loop voltage
control using the "inverted" machine model

123
Chapter 6. Power-HIL Testing of High-Power Electrical Machine Drives

analytically with the torque and voltage equations of the machine. "Inverted"
machine modeling has been used in [150, 154]. The main disadvantage of this
method is that only the fundamental current and voltage of the machine can
be emulated and not the current harmonics or the torque ripple. Low switching
frequency can be used for the emulator converter that allows the use of CPU-based
DSP instead of FPGA. Also, there is no need to measure the PWM ac voltage of
the inverter in order to implement this control. FEM data can still be utilized in
order to consider the variation of the stator inductance due to core saturation.

6.1.2 PMSM Emulation with the Inverted Machine Model


The "inverted" machine model has been studied further in this chapter, since it can
model accurately enough the dynamic and steady state response of a PMSM. The
detailed diagram of the developed P-HIL system and the structure of the control
for both converters is shown in Fig. 6.2. The IUT is controlled with standard
FOC, as described previously in Section 2.2. The control structure of the emulator
converter consists of three blocks: the machine model calculating the torque and
speed of the machine, the voltage reference calculation block and the modulator.
As can be seen in Fig. 6.3, input of the machine model block is the measured
three-phase current ie , which is converted into the dq0-frame. The calculated ie dq
is utilized to estimate the flux linkages ψd and ψq of the emulated machine through

(+)
Inverter under test Machine emulator
TA+ TB+ TC+ (+) Re Le TC+ TB+ TA+
Vdc + +
- -
Common
TA- TB- TC- TC- TB- TA-
mode choke
( not used here)
A
gate 6 gate limit applied
6
ie abc 010 110
SVM Voltage k at ma=0.75
010 110
ie abc Reference 011
111
000
100 (6.9)
Current 001 101
111
Calculation
011
000
100
controller θr SVM - +
θr ω r
001 101
ie dq with 0-seq. PI-
ie abc controller
Speed Machine model: compensation
Current reference - +
controller TL Torque/Speed
ωr generator i0
emulated Calculation
Look-up-table a
load torque 0A
b 1/3
+

Motor controller Real-time machine simulator

Fig. 6.2: Schematic of a PMSM emulator showing the circuit diagram of the topology
and the structure of the control system

124
6.2. Common Mode Voltage Control

look-up-tables that contain FEM-simulated data. Then, by substituting the


current and the flux linkage in (2.5), the electromagnetic torque Te is calculated.
Afterwards, by using (2.4), the acceleration dωr /dt can be derived which is
integrated to get the machine rotor speed ωr . With further integration of the rotor
speed, the rotor angle θr can also be calculated.
The voltage reference calculation block receives as input the calculated θr , ωr
and the machine current ie dq . The PMSM voltage equations shown in (2.7)-(2.8)
are modified in order to compensate for the voltage drop across the three-phase
inductor Le and its resistance Re . This voltage drop becomes significant especially
when high torque and speed are emulated.
dψd0
v∗em ≈ vem d − ωe ψq0
= (Rs − Re )ied + (6.1)
d
dt
dψq0
v∗em q ≈ vem q = (Rs − Re )ieq + + ωe ψd0 (6.2)
dt
where ωe = p ωr is the electrical frequency and p is the number of pole-pairs of
the machine. The stator copper loss of the machine is modelled through the stator
resistance Rs . The d- and q-axis flux linkages ψd0 , ψq0 are expressed as

ψd0 = ψd − Le id , ψq0 = ψq − Le iq (6.3)

The flux linkages ψd and ψq are again defined by 2-dimensional look-up-tables for
different d- and q-current combinations.
The final reference voltage of the emulator v∗em is sent to the SVM block that
generates the corresponding PWM signals.
ref. load
Flux Linkage torque T ωr Voltage *
L
dr Reference vem d
look-up-table
ie abc abc ie dq ψd Te θr θe
dt Calculation *

θe
dq ψq (2.5) (2.4)  ωr  p
(6.1)-(6.3)
vem q

ie dq

Fig. 6.3: Block diagram of the PMSM model as implemented in the "inverted" machine
emulator

6.2 Common Mode Voltage Control


Two-level converters have been used in this study and, therefore, the following
common mode voltage control is described for this topology.
SVM is used for both the IUT and the emulator VSC that has 8 available space
vectors, of which 6 are active vectors and 2 are zero vectors located at the origin

125
Chapter 6. Power-HIL Testing of High-Power Electrical Machine Drives

of the plane. The active vectors contribute mainly to the fundamental frequency
of the voltage, while the zero vectors create only zero sequence or else common
mode. The 2-level space vector hexagon is divided into 6 triangles and all the
three vectors of each triangles become active sequentially during each switching
period, as presented in Subsection 4.3.2.
All space vectors of the SVM plane contribute to both the Vαβ voltage components
and to the zero-sequence, as listed in Table 6.1. The values of these voltage
components originate from the Clarke transformation, when applied on the abc
phase voltage of each switching state. For example, the switching vector V 1 =
[1 0 0] creates the following phase voltages [Va Vb Vc ] = [ 12 Vdc − 12 Vdc − 12 Vdc ],
which can be transformed into the αβ-domain as Vα = 23 Vdc , Vβ = 0 and V0 =
− 16 Vdc . By multiplying these voltages with their corresponding duty cycles and
the dc-link voltage Vdc , the average output voltage of the inverter for one switching
period can be calculated.
An example is shown in Fig. 6.4(b) for reference voltage v̂∗ = 0.585 pu and
fundamental frequency 200 Hz. The duty cycle of phase A dphA is shown with
blue curve, while the duty cycles of the voltage vectors dE and dF are shown in red
and green, respectively. It can be observed that the duty cycles of both vectors
dE and dF have a dc and third-harmonic component that creates common mode
voltage equal to − 16 Vdc dE and 16 Vdc dF , respectively. The switching harmonics are
not considered with this calculation, since the above voltages are averaged within
one switching period.
The common mode voltage component of the long vectors V1 - V6 creates circulat-
ing current in the P-HIL test bench, as seen in the equivalent circuit of Fig. 6.5.
The two inverters can be represented by the common-mode-voltage sources Ve 0

Table 6.1: Switching Vectors of 2-level SVM [153]

Voltage Vector Vα /Vdc Vβ /Vdc V0 /Vdc


V 0 = [0 0 0] 0 0 -1/2
V 1 = [1 0 0] 2/3 0 -1/6

V 2 = [1 1 0] 1/3 1/ 3 1/6

V 3 = [0 1 0] -1/3 1/ 3 -1/6
V 4 = [0 1 1] -2/3 0 1/6

V 5 = [0 0 1] -1/3 −1/ 3 -1/6

V 6 = [1 0 1] 1/3 −1/ 3 1/6
V 7 = [1 1 1] 0 0 1/2

126
6.2. Common Mode Voltage Control

𝑽𝟑 𝟎𝟏𝟎 𝑽𝟐 𝟏𝟏𝟎
F
1
v̂*

Duty Cycle
𝑽𝟕 𝟏𝟏𝟏 D E
α
𝑽𝟒 𝟎𝟏𝟏 𝑽𝟎 𝟎𝟎𝟎 𝑽𝟏 𝟏𝟎𝟎 0.5 dphA
V dE
Vˆmax
*
 dc
3 dF
0
𝑽𝟓 𝟎𝟎𝟏 𝑽𝟔 𝟏𝟎𝟏 Vˆ *  0.5Vdc 0 0.005 0.01 0.015 0.02 0.025 0.03
Time [s]
(a) (b)

Fig. 6.4: 2-level space vector modulation indicating (a) the eight available switching
states, the reference voltage vector v̂∗ and with dashed circle the limit between
normal operation and overmodulation; (b) the duty cycles of phase A and of
the space vectors E and F for v̂∗ = 0.585 pu

dc Ve 0 ac ac Vem 0 dc
+ + Re /3 Le /3 + +
- 3ꞏi0 = (ie a + ie b + ie c)
-
- Rdc -
Lm
3_
Rdc i = (idc + + idc -)/2
Lm 2 0

Common mode choke


( not used here)
Fig. 6.5: Zero-sequence-equivalent circuit of the machine emulator with three-phase
inductor Le at the ac-side and common dc-link

and Vem 0 and the per-phase zero-sequence current is equal to


1
i0 = (ie a + ie b + ie c ) (6.4)
3
A common mode choke installed at the dc-link of the topology could filter the
common mode voltage and eliminate the circulating current. However, hardware
filtering is not preferred in this study. Instead, suitable control techniques are
studied in the following subsections that can suppress the circulating current
without any additional hardware.

6.2.1 Closed-loop Common Mode Current Control


The zero vectors V0 and V7 create only common mode voltage with opposite sign,
as seen in Table 6.1. In the standard SVM algorithm, the duty cycle dD from
(4.22) is usually distributed equally between these two vectors, as seen in (4.24), in

127
Chapter 6. Power-HIL Testing of High-Power Electrical Machine Drives

order to achieve average common mode voltage equal to zero during one switching
period. However, a different distribution of the zero vectors duty cycles can be
imposed in order to cancel out the the circulating current in the P-HIL test rig.
The circulating current is suppressed in the motor emulator of [152] through a
PI-controller that creates a reference common mode voltage v∗0 P I . This reference
voltage can adjust the distribution of the zero vectors in order to cancel the
circulating current. A similar approach has been followed in[157] to limit the zero
sequence current in parallel-connected 2-level inverters.
The circulating current dynamics depend on the series inductance of the circuit.
Therefore, the proportional and integral gains of this controller are selected as
kP 0seq = α0seq Le , kI 0seq = kP 0seq /3 (6.5)
where α0seq is the bandwidth of the control loop.
Since the purpose of the P-HIL set-up is to emulate the same operating conditions
for the IUT as if it was used with a real machine (i.e. same phase voltages
and currents), the standard SVM, as in (4.24), is used for the modulation of
that inverter. The common mode voltage is controlled only by the SVM of the
emulator-inverter.

6.2.2 Zero-Sequence Voltage Compensation


Although a PI control loop can regulate effectively the dc-component of the zero-
sequence, the bandwidth of the controller is not fast enough to eliminate completely
the third-harmonic component and, therefore, an improved control structure is
needed. More advanced control for the zero-sequence have been proposed for
parallel-connected inverters in [158] with a hybrid PWM technique and in [153]
presenting an instantaneous common mode voltage elimination technique using
SVM.
An improved control system is shown in Fig. 6.6 that manages to eliminate the
circulating current in machine emulators with 2-level inverters. The common
mode voltage created by the active voltage vectors within each switching cycle is
calculated and directly compensated by injecting equally sized voltage through
the zero vectors. Specifically, the injected zero-sequence voltage in per-unit can be
calculated as
1 1
v∗0 SV M = − dE + dF (6.6)
6 6
The PI regulator shown in the previous subsection with output v∗0 P I is also used
simultaneously in order to regulate the dc-component of the zero-sequence current

128
6.2. Common Mode Voltage Control

Inverter under test Machine emulator


SVM SVM
eq. eq. v* * eq.
d 0 SVM IUT v0 SVM em
010 110
010 110

(4.24) 1 d
(6.6) DEF 011 111
d2
011
111
000
100
(6.6) - +
000
100

001 101 001 101

eq.
k
eq. (6.7) (6.8) d1
d2
1
v0* PI
i0*  0 A
+
kI 0seq
 + x
iabc a - kP 0seq + /
b
c
+ 1/3
Vdc 0.5

Fig. 6.6: Schematic of the proposed zero-sequence-compensated space vector modulation

and ensure stability in the system. The total injected per-unit voltage for the zero
vectors of the emulator converter is
2
k = −v∗0 P I + (v∗0 SV M em − v∗0 SV M IU T ) (6.7)
dD
A limiter is applied on k with values of [-1,1] and the final duty cycles of the zero
vectors d1 and d2 are defined as
dD dD
d1 = (1 + k) , d2 = (1 − k) (6.8)
2 2

6.2.3 Operation close to the Voltage Limit


The proposed zero-sequence control works effectively when the reference voltage of
the emulator converter lays within V̂em∗
∈ [0, 0.5Vdc ], which is the area inside the
dotted circle in Fig. 6.4(a). For V̂em > 0.5Vdc the variable k exceeds the limits

and the duty cycle dD is not enough to fully compensate the total amount of
zero-sequence that flows in the circuit.
This is shown with the simulation results of a PMSM emulator in Fig. 6.7. Voltage
reference V̂em

= 0.49Vdc is used in the subplots (a), (c) and (e), whereas higher
voltage reference V̂em

= 0.53Vdc is used for the remaining (b), (d) and (f). The
variable k approaches the limits of [-1,1] in the first case, while it seems to get
saturated in the second case. While in the first case the circulating per-phase
current i0 remains under control, in the second case it becomes larger and almost
equal to the system where only a PI-controller regulates the common mode voltage.
In order to avoid such a condition that could cause instability in the common
mode voltage control of the inverter, the injected common mode voltage is reduced
at the high-voltage region as follows
2
k 0 = −v∗0 PI + c(v∗0 SV M em − v∗0 SV M IU T ) (6.9)
dD

129
Chapter 6. Power-HIL Testing of High-Power Electrical Machine Drives

1 1

0.5 0.5

0 0
k

k
-0.5 -0.5

-1 -1
800 801 802 803 804 805 806 1500 1501 1502 1503 1504 1505 1506
Time [ms] Time [ms]
(a) (b)

1 proposed SVM 1 proposed SVM


standard SVM standard SVM
Duty Cycle dphA

Duty Cycle dphA


0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
800 801 802 803 804 805 806 1500 1501 1502 1503 1504 1505 1506
Time [ms] Time [ms]
(c) (d)

40 40

20 20
i0 [A]

i0 [A]

0 0

-20 -20
proposed SVM proposed SVM
SVM with PI-control SVM with PI-control
-40 -40
800 801 802 803 804 805 806 1500 1501 1502 1503 1504 1505 1506
Time [ms] Time [ms]
(e) (f)

Fig. 6.7: Simulation results of PMSM emulator with zero-sequence voltage compensation
when V̂em

= 0.49Vdc in (a),(c),(e) and when V̂em

= 0.53Vdc in (b),(d),(f)

where

c = 1 , for V̂em

∈ [0, 0.5Vdc ]
√ ∗
1/ 3 − V̂em /Vdc Vdc (6.10)
c= √ , for V̂em

∈ (0.5Vdc , √ ]
1/ 3 − 0.5 3

130
6.3. Comparative Results of PMSM drive and Machine Emulator

6.3 Comparative Results of PMSM drive and Ma-


chine Emulator
A PMSM drive is tested in this section, both in transient and steady-state condi-
tions. The results of a standard machine test-bench are compared with similar
results from a machine emulator, which is tuned to emulate that PMSM.
Fig. 6.8 shows the laboratory set-up used for the PMSM test bench and for the
emulator of that machine. The same three-phase 2-level inverters are used for both
experiments, which are made with SiC power modules of the XM3 family [159].
The IUT has the CAB450M12XM3 modules, while the second inverter [marked as
"Load inverter" in Fig. 6.8(a) and "Machine emulator" in Fig. 6.8(b)], has the
CAB425M12XM3 modules. More details regarding the design of the SiC inverters
can be found in Chapter 7.
The inverters and the electrical machines are water cooled. A closed cooling loop is
built in the first set-up of Fig. 6.8(a) with a pump/ heat exchanger dissipating the
heat of the system, while running tap water is used for the set-up in Fig. 6.8(b).
The control system of the inverters is implemented on the dSPACE 1005 control
board.

DC dSPACE dSPACE
supply
IUT Load IUT
inverter DC
supply Le a

Le c

Le b AC
PMSM Load inductors
Machine
under PMSM
Heat emulator Le
test exchanger inverter

(a) (b)

Fig. 6.8: Laboratory set-up of (a) the 60 kW PMSM test bench and (b) emulator of this
PMSM drive using the same power converters

6.3.1 Experimental Results of a 60 kW PMSM Drive


The experimental results of the PMSM test bench of Fig. 6.8(a) are presented in
this subsection. The corresponding block diagram of the studied topology showing
the control of the inverters is shown in Fig. 6.9.

131
Chapter 6. Power-HIL Testing of High-Power Electrical Machine Drives

The two machines are identical and both controlled with FOC. Specifically, speed
control is applied on the machine-under-test where the input variable is the speed
reference ωr∗ . Torque control is used for the load machine with control input
being the reference torque Te2 ∗
. Look-up-table-based current reference generation
is applied that utilizes offline calculated data from FEM simulations. The optimal
stator currents Ie d and Ie q are shown in the machine torque-speed operation
map of Figs. 6.10(a)-(b), when the stator copper losses are minimized. The
resultant stator voltage and machine efficiency are shown in Figs. 6.10(c) and
(d), respectively. The mechanical losses of the efficiency subplot are calculated
by considering viscous damping equal to 3.8 · 10−4 Nm·s rad and the stray losses are
equal to 0.75% of the machine active power.
The parameters of the PMSM test-bench and the control variables of the two
inverters are listed in Table 6.2.
The dynamic response of the PMSM drive has been tested with the experimental
results in Fig. 6.11, showing that the controllers of the machines are stable and
follow their reference. The load machine operates in torque control mode as a
generator and the machine-under-test controls the rotational speed of the set-
up and operates as a motor. The speed of the system is set to 1000 rpm and,
then, at 1.194 sec the load torque increases in a ramp. Specifically, Fig. 6.11(a)
shows the reference ωr∗ and actual rotor speed ωr with blue color, the reference
electromagnetic torque for the load machine Te2 ∗
and for the machine-under-test

(+) Machine Load


Inverter under test Load Inverter
under test machine
TA+ TB+ TC+ (+) (+) TC+ TB+ TA+
Vdc + +
- -

TA- TB- TC- V V TC- TB- TA-


ve ac encoder ve2 ac
ve bc Tshaft [Nm] ve2 bc
A A
gate 6 gate
SVM 6
* ie2 abc
010 110
ve  ie abc angle ve*2  010 110

111
Current θr Current 011
111
000
100
011 100
controller controller
 vmax
000

 vmax
001 101
001 101


ie* dq ie*2 dq SVM

Current reference Te* Speed Current reference


generator controller r*
ωr ωr generator
Look-up-table
Look-up-table Load-side
ωr
Motor controller Te*2 Controller

Fig. 6.9: Schematic of a PMSM test bench showing the circuit diagram of the topology
and the structure of the control system

132
6.3. Comparative Results of PMSM drive and Machine Emulator

(a) (b)

(c) (d)
Fig. 6.10: Simulation results of the (a) d-axis current, (b) q-axis current, (c) stator
terminal voltage amplitude and (d) efficiency map when stator copper loss is
applied on the tested PMSM

Table 6.2: PMSM Rated Parameters & Control

Parameter Value Unit


Rated active power 30 kW
Peak active power 60 kW
Measured viscous damping breal 8.93·10−4 Nm·s/rad
Static friction T0 0.96 Nm
Shaft inertia J 0.047 kg·m2
Number of pole pairs p 4 -
Stator resistance Rs 26.4 mΩ
No-load stator inductance Ld 0.44 mH
No-load stator inductance Lq 1.3 mH
Magnet flux linkage ψmd 69.4 mWb
Inverter dc voltage Vdc 300 V
Inverter switching freq. fsw 20 kHz
Current control bandwidth ac 1000 rads/s
Speed control bandwidth aω 20 rads/s
Viscous damping (used only in the
0.02 Nm·s/rad
speed controller gains) b

133
Chapter 6. Power-HIL Testing of High-Power Electrical Machine Drives

60 100
ie d [A]

Stator currents ie [A]


1000
Rotor speed [rpm]

ie q [A]

Torque [Nm]
800 40 50
600
20
400 0
200 0
X 1.194 * *
Te2 Te
*
Y0 r r
0 -50
0 XY 0.1945
0
2 4 6 8 10 0 2 4 6 8 10
Time [s] Time [s]
(a) (b)

Fig. 6.11: Ramp test of the PMSM drive with speed set to 1000 rpm and reference load
torque Te2

= 40 Nm

Te∗ with red color. The torque Te2 ∗


is set to 40 Nm, while Te∗ is the output of
the speed control loop and is equal to 53 Nm, because it supplies also the losses
of the machine bench. Fig. 6.11(b) shows the measured stator currents of the
machine-under-test ie d and ie q .
Steady-state test results of the PMSM drive are shown in the following Fig. 6.12
and in Table 6.3. The input power of both machines is measured using two high-
bandwidth voltage probes and two current probes at terminals of each machine for
measuring with the oscilloscope the line-to-line voltages and currents. Sampling
frequency of 2.5 MHz is used. Then, FFT is performed on the measured signals
and the average power of the machines is calculated using the fundamental current
and voltage.
Experimental tests with speed higher than 3000 rpm have not been possible to be
conducted with the current laboratory set-up due to mechanical limitations of the
machine test-bench.

400 400
Stator voltage ve2 ab [V]
Stator voltage ve ab [V]

Stator current ie2 a [A]


Stator Current ie a [A]

200 200
200 200
100 100

0 0 0 0

-100 -100
-200 -200
-200 -200
-400 -400
0 2 4 6 0 2 4 6
Time [s] 10-3 Time [s] 10-3

(a) (b)

Fig. 6.12: Measured stator line voltage ve ab and phase current ie a from the (a) PMSM
under test and (b) the load machine with ωr = 2500 rpm and Te2 ∗
= 100 Nm

134
6.3. Comparative Results of PMSM drive and Machine Emulator

Table 6.3: PMSM Drive Steady-State Experimental Results

Reference ωr ωr ωr ωr ωr
load 1000 2500 1000 2500 1000 2500 1000 2500 1000 2500
torque rpm rpm rpm rpm rpm

Te2 Te∗ [Nm] Iˆe (1) [A] V̂e (1) [V] Pe1 [kW] Pe2 [kW]
[Nm] ref. torque fundamental fundamental input power input power
PMSM 1 ph. current ampl. volt. ampl. PMSM 1 PMSM 2
0 1.01 5.29 3.17 13.68 27.30 66.46 0.1 1.17 0 0.01
20 25.94 27.41 53.46 55.98 28.84 66.48 2.29 5.49 -1.89 -4.90
40 50.80 54.24 93.24 97.57 34.39 76.51 4.75 11.11 -3.80 -10.07
60 76.07 79.30 129.63 133.22 38.86 87.41 7.25 16.67 -5.47 -14.92
80 98.88 103.45 162.41 167.54 43.77 97.43 9.87 22.33 -7.14 -19.71
100 121.24 125.23 195.15 198.97 47.88 108.12 12.53 28.04 -8.76 -24.56

6.3.2 Experimental Results of PMSM Emulator


The PMSM that has been tested in the previous subsection is now emulated on
the P-HIL set-up of Fig. 6.8(b). The control diagram of the system can be seen in
Fig. 6.2. The flux linkage maps of the machine obtained through FEM simulation
of the PMSM, which are shown below in Fig. 6.13 as a function of the stator
currents, are utilized in order to calculate the terminal voltage of the emulated
machine. The same control parameters are used for the IUT, as with the PMSM
drive previously reported in Table 6.2.
A three-phase inductor is installed at the ac-side of the topology with inductance
Le = 0.55 mH and resistance Re = 53 mΩ. No common mode hardware filter has
been installed in the system, since the circulating zero-sequence current is regulated

(a) (b)

Fig. 6.13: Flux linkage maps of the emulated PMSM

135
Chapter 6. Power-HIL Testing of High-Power Electrical Machine Drives

effectively with the common mode voltage control presented in Subsection 6.2.2.

PMSM Model Evaluation:


Accurate estimation of the machine mechanical and electrical parameters is impor-
tant in order to emulate accurately the currents, voltages, speed and torque of the
PMSM. Therefore, the general form of the mechanical dynamics equation (2.4)
has been used here with static friction torque T0 = 0.96 Nm, viscous damping
breal = 8.93 · 10−4 Nm·s/rad and inertia J = 0.047 kg·m2 . These parameters
have been measured experimentally through dynamic response tests of the PMSM
test-bench.
The fidelity of the PMSM emulator is evaluated with the following steady-state
experimental results in comparison with the results of the previous subsection,
where the real PMSM drive was tested. Due to the presence of common mode
voltage and current, all the three phase voltages and line currents need to be
measured with the oscilloscope in order to calculate accurately the power at the
ac output of the IUT Pe1 .
Z T1 Z T1
1 1
Pe1 = pe1 (t) dt = [veaN (t)iea (t) + vebN (t)ieb (t) + vecN (t)iec (t)] dt
T1 0 T1 0
(6.11)
where pe1 (t) is the instantaneous active power and T1 the fundamental period.
Time-domain results are shown in Fig. 6.14 for ωr = 2500 rpm and reference
torque Te∗ = 100.43 Nm (for the emulator) and Te∗ = 103.45 Nm (for the real
PMSM). More results are listed in Table 6.4.
It should be noted here that the load torque of the drive (first column of the
table) is controlled differently in the PMSM emulator compared to the real PMSM
and, therefore, these values do not correspond to the same operating point. In
the emulator the TL is a controllable input of the machine model (see Fig. 6.3)
and the reference torque of the IUT Te∗ is the output of the speed control loop
and is the sum of the load torque and the mechanical losses, represented by the
viscous damping breal and static friction T0 . As for the real PMSM drive, the load
torque is regulated indirectly through the reference torque of the load machine Te2

.
Therefore, the reference torque of the PMSM-under-test Te is the sum of the Te2 ,
∗ ∗

the losses of the load machine, the losses of the two inverters and the core losses
of the PMSM-under-test. These losses have not been considered in the model of
the machine emulator and a difference can be seen between the values of Te∗ for
the same load torque in Table 6.3 and Table 6.4.
The PMSM torque would have been calculated more accurately, if the torque on

136
6.3. Comparative Results of PMSM drive and Machine Emulator

(a) (b)
60 60

Reactive power [kVAr]


Active power [kW]

40 40

20 20

0 0
Drive pe1 Emulator pe1 Drive Pe1 Emulator Pe1 Drive qe1 Emulator qe1 Drive Qe1 Emulator Qe1

0 1 2 3 4 5 6 0 1 2 3 4 5 6
Time [s] 10-3 Time [s] 10-3
(c) (d)

Fig. 6.14: Comparative results of the PMSM drive and emulator with ωr = 2500 rpm
and reference torque Te∗ = 100.43 Nm (for the emulator) and Te∗ = 103.45 Nm
(for the real PMSM drive)

Table 6.4: PMSM Emulator Steady-State Experimental Results

Emulated ωr ωr ωr ωr
load 1000 2500 1000 2500 1000 2500 1000 2500
torque rpm rpm rpm rpm
TL Te∗ [Nm] Iˆe (1) [A] V̂e (1) [V] Pe1 [kW]
[Nm] ref. torque fundamental fundamental input power
emulated PMSM ph. current ampl. volt. ampl. emulated PMSM
0 0.76 0.80 2.42 2.71 31.49 75.26 0.11 0.30
20 21.65 21.66 45.23 45.36 39.28 89.35 2.29 5.94
40 40.50 40.48 78.57 78.64 46.47 107.52 4.75 11.55
60 60.46 60.43 108.26 108.43 51.03 119.27 7.25 17.03
80 80.43 80.42 137.10 137.44 54.18 127.40 9.87 22.44
100 100.43 100.43 166.22 166.69 56.46 132.17 12.53 27.76

the shaft of the machine Tshaf t had been measured with a torque sensor. Then,
the measured Tshaf t would be similar with the load torque of the emulator TL .
However, this has not been possible during this project.

137
Chapter 6. Power-HIL Testing of High-Power Electrical Machine Drives

When operating points with the same reference torque Te∗ are compared, the
current of the IUT in the machine emulator is almost identical with the real
PMSM current. However, there is a discrepancy between the emulator and the
drive voltage, with the first one being slightly larger. This can be attributed to
error in the compensation of the voltage drop across the inductor Le and also to
wrong estimation of the machine flux linkage, since FEM data are utilized which
vary from the real PMSM flux linkage. For example for the plotted operating
point in Fig. 6.14, the fundamental current amplitude is 166.69 A for the emulator
and 167.54 A for the drive (a difference of 0.5%), whereas the fundamental voltage
amplitude is 132.17 for the emulator and 97.43 V for the drive. More accurate
estimation of the inductance value Le and measurements of the flux linkage of the
PMSM could reduce this error in the voltage.

Common Mode Voltage Control Evaluation:


The proposed common mode voltage control of the machine emulator is validated
with experimental results at various operating points of the emulated PMSM. Two
versions of the machine emulator are tested. One with just a PI-controller-based
zero-sequence regulator (as shown in Subsection 6.2.1), named as "SVM1" in the
following results. The second version of the system has the proposed zero-sequence
compensated SVM (as shown in Subsection 6.2.2) and is named "SVM2". There
is no common mode hardware filter in any of the systems examined.
The PI-regulator of the common mode voltage controller in both cases has band-
width equal to a0seq = 10ac = 10000 rad/s. The individual gains are selected
based on (6.5).
Firstly, the response of the emulator is tested without any common mode voltage
control and the results are shown in Figs. 6.15(a)-(b) for ωr = 500 rpm / Te∗ =
41 Nm and in Figs. 6.15(c)-(d) for ωr = 3000 rpm / Te∗ = 61 Nm. The zero-
sequence current is proportional with the voltage difference between the IUT and
the emulator converter (Ve0 − Vem0 ), since only the equivalent resistance of the
resultant loop limits the current. Therefore, the second tested case with the higher
emulated speed has also higher i0 , since the power of the system and the inverter
voltage in the second case are higher. Based on these results it is clear that
controlling the common mode voltage in order to reduce the circulating current is
necessary in order to have safe operation of the machine emulator.
The performance of the two systems with "SVM1" and "SVM2" is evaluated with
the comparative experimental results of Figs. 6.16-6.18. The response of the
system during one fundamental period at three operating points is demonstrated.

138
6.3. Comparative Results of PMSM drive and Machine Emulator

Zero-sequence current [A]


100 i0
Stator current ie [A]

t 0.01115
20 i0 14.27
I0
50 10
t 0.01466
0 0 I0 -3.519

-50 -10
Phase A
Phase B -20
-100 Phase C
0 0.005 0.01 0.015 0.02 0.025 0.03 0 0.005 0.01 0.015 0.02 0.025 0.03
Time [s] Time [s]
(a) (b)
t 0.003608

Zero-sequence current [A]


i0 22.4
100
Stator current ie [A]

20
50 10 t 0.002465
i0 1.415

0 0

-50 -10
Phase A i0
Phase B -20
-100 Phase C I0

0 1 2 3 4 5 0 1 2 3 4 5
Time [s] 10-3 Time [s] 10-3
(c) (d)

Fig. 6.15: PMSM emulator phase current ie a and zero-sequence current i0 with no control
of the common mode voltage when (a)-(b) ωr = 500 rpm / Te∗ = 41 Nm and
(c)-(d) ωr = 3000 rpm / Te∗ = 61 Nm
Zero-sequence current [A]

t 0.001786
150 i0 19.07
SVM1
Stator current ie a [A]

20
100 SVM2
t 0.002273
10 i0 4.133
50 t 0.003188
I0 0.000708

0 0

-50 -10

-100 -20
SVM1 i0 I0 SVM2 i0 I0
-150 -30
0 1 2 3 4 5 0 1 2 3 4 5
(a) emulator phase current ie a 10-3
TimeA[s] Time [s]i0 and average
(b) zero-seq. instantaneous I0
10-3
1 1
Standard SVM
Duty cycle dphase A

0.8 SVM1
SVM2 0.5
0.6
0
k

0.4
-0.5
0.2 SVM1
SVM2
0 -1
0 1 2 3 4 5 0 1 2 3 4 5
Time [s] 10-3 Time [s] 10-3
(c) duty cycle of the emulator converter (d) common mode ref. voltage k (in per-unit)

Fig. 6.16: Comparative results of the "SVM1" and "SVM2" when ωr = 3000 rpm /
Te∗ = 61 Nm and modulation index of emulator inverter ma em = 0.66

139
Chapter 6. Power-HIL Testing of High-Power Electrical Machine Drives

Zero-sequence current [A]


150
SVM1
Stator current ie a [A]
20 t 0.0005844
i0 22.47
100 SVM2
t 0.001098
10 i0 5.2
50
0 0 t 0.002793
I0 0.2412

-50 -10

-100 -20
SVM1 i0 I0 SVM2 i0 I0
-150 -30
0 1 2 3 4 5 0 1 2 3 4 5
(a) emulator phase A[s]current ie a 10-3
Time Time [s]i0 and average
(b) zero-seq. instantaneous I0
10-3
1 1
Standard SVM
Duty cycle dphase A

0.8 SVM1
SVM2 0.5
0.6
0

k
0.4
-0.5
0.2 SVM1
SVM2
0 -1
0 1 2 3 4 5 0 1 2 3 4 5
Time [s] 10-3 Time [s] 10-3
(c) duty cycle of the emulator converter (d) common mode ref. voltage k (in per-unit)

Fig. 6.17: Comparative results of the "SVM1" and "SVM2" when ωr = 3000 rpm /
Te∗ = 81 Nm and modulation index of emulator inverter ma em = 0.69
Zero-sequence current [A]

150 t 0.001724
SVM1
Stator current ie a [A]

20 i0 14.73
100 SVM2
10 t 0.002092
50 i0 3.2 t 0.002685
I0 0.1933
0
0
-50 -10

-100 -20
SVM1 i0 I0 SVM2 i0 I0
-150 -30
0 0.5 1 1.5 2 2.5 3 3.5 0 0.5 1 1.5 2 2.5 3 3.5
(a) emulator phase
TimeA[s]
current ie a 10-3 Time [s]i0 and
(b) zero-seq. instantaneous average I0
10-3
1 1
Standard SVM
Duty cycle dphase A

0.8 SVM1
SVM2 0.5
0.6
0
k

0.4
-0.5
0.2 SVM1
SVM2
0 -1
0 1 2 3 0 0.5 1 1.5 2 2.5 3 3.5
Time [s] 10-3 Time [s] 10-3
(c) duty cycle of the emulator converter (d) common mode ref. voltage k (in per-unit)

Fig. 6.18: Comparative results of the "SVM1" and "SVM2" when ωr = 4000 rpm /
Te∗ = 31 Nm and modulation index of emulator inverter ma em = 0.79

At ωr = 3000 rpm / Te∗ = 61 Nm / ma em = 0.66 (Fig. 6.16) and ωr = 3000 rpm /


Te∗ = 81 Nm / ma em = 0.69 (Fig. 6.17) the emulated machine speed is lower than
the base speed. Fig. 6.18 shows the results for ωr = 4000 rpm / Te∗ = 50 Nm /

140
6.3. Comparative Results of PMSM drive and Machine Emulator

ma em = 0.79, which belongs to the high-speed operating region of the PMSM. In


all three cases the control scheme "SVM2" performs better and keeps the current
i0 much lower than the "SVM1".
Specifically in the first tested case of Fig. 6.16(b), the peak of i0 equals 3.7% of the
phase current amplitude for the "SVM2", while it is equal to 16.7% of the phase
current amplitude for the "SVM1", which is a 78% reduction. This results to more
sinusoidal current ie a for the "SVM2", which is closer to the stator current of the
real PMSM. A large amount of third-harmonic component is present in the current
when using the "SVM1" controller. The common mode reference voltage k in Fig.
6.16(d) has similar amplitude for both control schemes. However there seems to be
a phase difference between the two variables, which is the reason why the PI-based
"SVM1" is not as effective as the "SVM2". The "SVM2" generates the correct
amplitude and angle for the common mode voltage, so that the zero-sequence
current can be suppressed.
The k generated by "SVM2" starts to have larger amplitude than the one generated
by "SVM1" and approaches the limits of [−1, 1] in Figs. 6.17(d) and 6.18(d) as
the modulation index of the emulator inverter increases, which complies with the
simulation results of Subsection 6.2.3.
Further experimental results are shown in Fig. 6.19 of operating points with even
higher modulation index for the emulator converter. As seen in Fig. 6.19(b),(c)
the controller reference voltage k gets saturated at ±1, since the proposed common-
mode-voltage-compensation scheme "SVM2" approaches its theoretical limit (based
on Subsection 6.2.3). However, the zero-sequence current still remains low, with
peak values 9.1 A, 6.3 A and 3.3 A (3.3%, 3.8% and 7% of the ac current amplitude,
respectively).
It has not been possible to test operating points on the PMSM operating map
with even higher voltage, since the voltage limit of the IUT had already been
reached in Fig. 6.19, as seen by the modulation index of the IUT ma e . However,
it is expected that the common mode voltage controller of "SVM2" would be even
more saturated and the resultant i0 would gradually increase, if ma em closer to 1
is used.

141
Chapter 6. Power-HIL Testing of High-Power Electrical Machine Drives

t 0.003342

Zero-sequence current [A]


i0 267.8
300 20 1
Stator current ie a [A]
t 0.0002735
200 ie a 9.067
10 0.5
100
0 0 0

k
-100
-10 -0.5
-200
-300 -20 -1
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Time [s] 10-3 Time [s] 10-3

(a) ωr = 2500 rpm / Te∗ = 161 Nm / ma em = 0.64 / ma e = 0.84

Zero-sequence current [A]


200 20 1
Stator current ie a [A]

100 10 0.5

0 0 0

k
-100 -10 -0.5

-200 -20 -1
0 1 2 3 4 5 0 1 2 3 4 5
-3
Time [s] 10 Time [s] 10-3

(b) ωr = 3000 rpm / Te∗ = 101 Nm / ma em = 0.72 / ma e = 0.94


Zero-sequence current [A]

200 20 1
Stator current ie a [A]

100 10 0.5

0 0 0
k

-100 -10 -0.5

-200 -20 -1
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
Time [s] 10-3 Time [s] 10-3

(c) ωr = 4500 rpm / Te∗ = 21 Nm / ma em = 0.84 / ma e = 0.93


Fig. 6.19: Results of the PMSM emulator with "SVM2" as the common mode voltage
control scheme

6.4 Summary
This chapter has presented the design process and control of a P-HIL set-up with
2-level inverters for emulating the operation of a PMSM. The P-HIL set-up consists
of two identical converter, named as inverter-under-test (IUT) and emulator VSC,
that share the same dc-link and are connected also at the ac side through a
three-phase inductor.
The "inverted" machine modelling approach has been followed, where the machine
is modelled with its analytical voltage, speed and torque equations. The machine
response is modelled by the resultant voltage which is generated by the emulator

142
6.4. Summary

converter. The modelling process of the PMSM has been explained and experi-
mental results on a test-cases PMSM emulator are shown in comparison with a
real PMSM drive. The current of the PMSM for each individual torque/speed
point is emulated with high accuracy with error as low as 0.5%. However, the
terminal voltage of the PMSM is emulated with larger error, which is attributed
to inaccuracies of the imported machine parameters and the voltage drop across
the inductor that exists in the P-HIL set-up.
A control technique is also presented for suppressing the zero-sequence current
that naturally flows in this P-HIL topology. The proposed controller consist of a
combination of a PI-regulator and on-the-fly compensation of the common mode
voltage in the SVM block of the emulator converter. Experimental results have
verified the effectiveness of the proposed control, which is also compared with a
simpler common mode voltage controller consisting only of a PI-regulator. The
proposed control scheme limits the zero-sequence current amplitude to 3.3% of the
main phase current amplitude, when active power of the PMSM equal to 42 kW
and its base speed are emulated.

143
Chapter 6. Power-HIL Testing of High-Power Electrical Machine Drives

144
Chapter 7

Electrical and Thermal Design of


Automotive SiC Inverter

The design process of a three-phase, 2-level inverter intended to be used in


automotive applications is studied in this chapter. The design specifications of
motor drives for electric vehicles are firstly analyzed. Then, a prototype SiC
inverter is manufactured based on these specifications. This inverter has been
used with the experimental set-up of the previous Chapter 6.

7.1 Design Specifications


The electric drive, including the motor and its inverter, is one of the core compo-
nents of an electric vehicle. Electrification of ground vehicle transportation has
been a fast-growing research area and market in the recent years and the electric
drive is one of the main research points. Correct dimensioning of the drive and
selection of its components is of high importance in order to meet effectively the
design requirements of the application.
The design specifications vary depending on the application where the vehicle is
being used. The basic classification of vehicles based on their mass is between
light- (≤ 3855 kg) and heavy-duty vehicles (> 3855 kg), according to EPA [160].
The same distinctions exist also from other governmental authorities, such as in
[161] for the EU. Passenger vehicles, vans and minivans belong to the light-duty
vehicles, while trucks, buses and coaches are heavy-duty vehicles. Off-road vehicles
is another special classification.

145
Chapter 7. Electrical and Thermal Design of Automotive SiC Inverter

7.1.1 Design Specifications for Light-Duty Vehicles


A detailed analysis of the legislation and the design assessment of drives for light-
duty vehicles can be found in [162]. The first step in the design of the drive is the
modelling of the vehicle’s motion, such as the forces applied on the wheels and
the acceleration. The acceleration avehicle is described as
dvvehicle
m avehicle = m = Facc (7.1)
dt
where m [kg] is the equivalent mass of the moving object. The acceleration force
is calculated as
Tmotor
Facc = gear ngear − Froad load (7.2)
r
where gear and ngear are the gear ratio and efficiency of the transmission system,
respectively. The wheel radius is defined as r and the load of the wheels is Froad load .
The Froad load consists of the air drag Fair , the rolling resistance Froll and the
grading force of the road Fgrade . These are described as
1
Fair = ρair Cd A (vvehicle − vwind )2 (7.3)
2
Froll = Cr m g cos(α) (7.4)
Fgrade = m g sin(α) (7.5)

where vvehicle and vwind [m/s] are the forward speed of the vehicle and the wind,
%road_slope
g [m/s2 ] is the gravitational acceleration and α = tan−1 ( 100 ) [rad] is the
road angle of inclination. Cd is the aerodynamic drag, Cr the rolling resistance
and A [m2 ] the effective cross sectional vehicle area, which are parameters specific
for each vehicle. The parameter A depends on the frontal area of the vehicle and
for this study it is set equal to A = 86% · height · width, as referenced from [162].
There are many testing procedures suitable for light-duty vehicles, as described
in [163–165]. The vehicle-under-test, modelled with (7.1)-(7.5), can be simulated
when these tests are applied. The acceleration test determines the maximum
acceleration and the maximum speed that can be maintained by the electric
drive of the vehicle for different road slopes. Similarly, deceleration tests would
determine the response of the vehicle to the road load force.
Standardized tests with legislative driving cycles is another method of examining
the vehicle operation, when it operates with a predefined speed profile. These
cycles represent the average driving pattern of the users and traditionally they
have been used to calculate the fuel consumption and emissions of non-electric
vehicles. In the case of electric vehicles, these cycles can be utilized to simulate the

146
7.1. Design Specifications

100
low speed medium speed high speed low speed high speed

Vehicle speed [km/h]


Vehicle speed [km/h]
100
80
80
60
60
40
40
20 20

0 0
0 200 400 600 800 1000 1200 1400 1600 1800 0 200 400 600 800 1000 1200 1400 1600 1800
Time [s] Time [s]
(a) (b)

Fig. 7.1: China automotive test cycles (a) for passenger vehicles (CLTC_P) and for (a)
for large heavy-duty vehicles (CHTC_TT)

load of the electric drive, as well as the aging of these components. There are many
cycles suitable for light-duty vehicles, such as the Worldwide Harmonised Light
Vehicles Test Cyle (WLTC) [166], which is currently the main drive test procedure
in many countries, the New European Driving Cycle (NEDC), the ARTEMIS [167],
etc. A new standard has also been developed by the China-Automotive-Test-Cycle
program, and has been implemented in China in May 2020. New driving cycles are
introduced by this standard for each vehicle category and the one corresponding
to the passenger vehicles is named CLTC_P [168]. The speed profile of that cycle
is shown in Fig. 7.1(a) and will be further studied in this chapter.

7.1.2 Design Specifications for Heavy-Duty Vehicles


There are some distinctive differences in the requirements of the electric drive
design for heavy-duty vehicles, like long-haul trucks, compared to light passenger
vehicles. A high peak torque is required in the low-speed operating region in order
to startup the vehicle [169]. A wide speed range with high power at the flux
weakening region is needed [170], since the drive operates frequently and for a long
duration close to its rated power, especially when the truck is driven on a road
with large uphill slope or on the highway with a constant speed close to its limit.
The most common operating conditions of a truck are summarized in Table
7.1.Firstly, the start-up conditions of the drive require a high peak torque in order
to be able to accelerate the vehicle from standstill while being on a road with slope.
The "high power operation" corresponds to driving on an uphill road and the
"best efficiency operation" is the high-speed driving condition on a highway, which
is the most common for long-haul trucks. The numbers in this table are given as
an example for a 40 tonnes vehicle, with peak power 500 kW and maximum speed

147
Chapter 7. Electrical and Thermal Design of Automotive SiC Inverter

Table 7.1: Test-Case Truck Electric Drive Operation Requirements

Parameter Value Unit


(1) Start-up Operation with Peak Torque
Parameter Value Unit
Duration 5 min
(3) Best Efficiency Operation
Road slope 12 %
Duration for a long time
(2) High Power Operation
Power 100 kW
Duration 20 min
Road slope 0 %
Power 400 kW
Vehicle speed 80 km/h
Road slope 6 %
Vehicle speed 50 km/h

100 km/h. This test-case has been studied with more details in 1
.
Typical driving conditions of large trucks can also be modelled with suitable
driving cycles, such as the CHTC_TT, shown in Fig. 7.1(b).

7.2 SiC 2-Level Inverter Design Requirements


Two SiC 2-level inverters with identical layout have been designed in this project
satisfying the previously mentioned specifications of electric vehicles in terms of
power capability. Based on current industry trends in motor drives for BEVs,
800 V rated dc voltage and approximately 200 kW rated active power (250 kW
peak) of the prototypes have been selected. The 2019-released SiC power modules
of the XM3 half-bridge family [159], manufactured by Wolfspeed, satisfy these
requirements and at the same time they offer small form-factor and low parasitic
inductance enabling high-speed switching with low voltage ringing.
The CAB450M12XM3 module has been used in one of the prototypes (named as
"IUT" in Chapter 6) that is optimized for low-conduction losses with a switching
frequency of 10 to 20 kHz, typical values for automotive drives. The second
prototype has CAB425M12XM3 modules, which are more optimized towards high
switching frequency and is used as the "machine emulator inverter" in Chapter 6.
The dc-link capacitors of the inverter are responsible for filtering out the current
ripple caused by the PWM switching pattern, so that the dc-side source can supply
1 G. Mademlis, Y. Liu, J. Tang, L. Boscaglia, and N. Sharma, "Performance Evaluation of

Electrically Excited Synchronous Machine compared to PMSM for High-Power Traction Drives",
in Proc. 24th International Conference on Electrical Machines (ICEM), August 23-26, 2020, pp.
1793-1799.

148
7.2. SiC 2-Level Inverter Design Requirements

clean current without any harmonics. Correct sizing of the dc-link capacitor bank
in terms of current ripple capability, voltage rating and capacitance is important
in order to avoid overheating of the capacitors and to filter effectively all the
harmonics. Based on the literature [171–173], film capacitors are the optimal
choice for automotive applications compared to electrolytic ones due to their lower
ESR ratings, leading to much better handling of the fast-switching current ripple
and higher inverter efficiency. They also have smaller size and higher efficiency,
due to their self-healing capability. Ceramic capacitors can also be a candidate for
even lower ESR and better high-frequency capabilities [174, 175], however, with a
much higher cost than film capacitors.
The rms ac current of a three-phase inverter, modulated with SVM, having dc-link
voltage Vdc , line-to-line rms voltage Ve rms and ac-side power Pe can be calculated
as
Pe (4.8) Pe
Ie rms = √ == √ (7.6)
3 ninv P F Ve rms 3 ninv P F 0.98V
√ dc
2

where P F is the power factor, ninv the inverter efficiency and 0.98 is the a
maximum allowed modulation index.
The current stress of the dc-link capacitors depends on the power factor and the
power rating of the inverter. The worst-case capacitor current Ic rms for inverters
controlling a PMSM can be estimated as follows [176]
1
Ic rms ≈ √ Ie rms (7.7)
2
Considering Pe = 250 kW, ninv = 0.93, P F = 0.93 and Vdc = 800 V, the capacitor
current requirement of the inverter is Ic rms = 213 A.
The required capacitance of the dc-link is determined by the allowed voltage ripple
Ts
P
Ts ∆ic ripple
t=0
Cdc min = (7.8)
∆Vdc max
where ∆Vdc max = 5%Vdc is the maximum ripple of the dc voltage. The resultant
lowest required capacitance is equal to Cdc min = 77 µF for switching frequency
20 kHz.
Considering the above calculations and the stock availability at the time when
the prototype was built, automotive-graded film capacitors from KEMET with
product code C4AQQBW5400A3NJ were selected. These capacitors have ratings
1.1 kV, 40 µF, Ic rms,10kHz = 18 A and RESR,10kHz = 4 mΩ. In order to satisfy

149
Chapter 7. Electrical and Thermal Design of Automotive SiC Inverter

the current requirement of 213 A, 12 pieces of this capacitor connected in parallel


are needed.
As can be observed by the rated parameters of the selected product, the equivalent
capacitance of the inverter is over-dimensioned, since 12 · 40µF >> Cdc min . This
shows that the main design limitation when selecting film capacitors for low-voltage
and high-current inverters is the capacitor current. As highlighted in [177], the
current-carrying capability of film capacitors needs to be further increased by
future research, in order to allow the design of SiC inverters with even higher
power density. On the other hand, it is easier to satisfy the capacitor current
specifications of higher-voltage inverters, such as with the NPC inverter prototype
of Subsection 5.2.3.

7.3 DC-Link Capacitor Bank Design


The 12 selected film capacitors need to be connected to the dc terminals of the
power modules and in parallel with each other. It is important the capacitors to
be connected with the lowest possible parasitic resistance and inductance (ESR
and ESL) in order to avoid voltage ringing. A special design of 3 mm-thick copper
bars is made, connected to the DC+ and DC- terminals, so that the round pins
of the capacitors can be soldered there. FEM simulations of the copper bars
are conducted, as seen in Fig. 7.2(a), in order to calculate the resultant current
density and to make sure that areas with too high current density are avoided.

(a) current density distribution in A/mm2 (b)

Fig. 7.2: DC-link copper bars (a) FEM simulated with COMSOL when 400 A dc current
is applied between the left and the right side of the bars and (b) the capacitor
bank after being soldered

150
7.3. DC-Link Capacitor Bank Design

High current density is just observed on the right side of the bars, close to the
dc-link terminals.
It is also possible to use capacitor banks that are pre-assembled in brick structures,
as seen in [178]. However, there has not been any suitable brick capacitor available
that could fit the terminals of the specific SiC power modules. A different approach
with a combination of aluminium plates and PCB for soldering the capacitors
has been presented in [174]. However, such designs are not suitable for the power
ratings of this project.
It is important to solder the capacitors with high-enough temperature, so that a
proper joint is built between the soldering material and the copper. Cold-soldering
should be avoided, since it would cause additional contact resistance increasing
the ESR of the capacitors and the joint can also break easily. Many capacitor pins
seem not to be properly soldered on the copper busbars of Fig. 7.2(b), which are
reheated and fixed at a later stage. Standard soldering material with Sn-Pb alloys
melts at around 180-190°C. Solder wire with different materials that have lower
melting point, such as indium-based alloys, can also be used.
The equivalent parallel capacitance and impedance of the final version of the
distributed capacitor bank is measured with an LCR meter and the results are
shown in Fig. 7.3 for a wide frequency range. The measured capacitance at dc
frequency is 471.6 µF, which is close to the theoretical value of 480 µF. The
impedance waveform is typical for film capacitor banks [179]. In low-frequency
regions the measured impedance decreases slowly and inversely with frequency up
to the resonance point of 408 kHz, where it reaches a minimum impedance value
| Zmin |= RESR = 6.5 mΩ. For frequencies higher than the resonant frequency,
the capacitor bank has inductive performance.

f 0.253
0.5 C 0.4716 6
Capacitance [mF]

0.4
4
0.3
Impedance |

0.2
2
0.1

0 0 f 407900
0 50 100 150 200 103 104 105 | | 0.0645 106 107
Frequency [kHz] Frequency (log) [Hz]
(a) (b)

Fig. 7.3: Measurements with LCR meter of the equivalent (a) capacitance and (b)
impedance versus applied frequency

151
Chapter 7. Electrical and Thermal Design of Automotive SiC Inverter

The final assembled inverter is shown in Fig. 7.4, when mounted inside a metal
enclosure for better EMI shielding. A current sensor PCB has also been designed
and mounted at the ac side of the inverter, having the Hall-effect sensors ML91208
from Melexis. The main advantage of these sensors is their small footprint; however,
they are susceptible to EMI and have lower accuracy than other closed-loop sensors.

Gate drivers
(& CAB425M12XM3 PWM
beneath)
signals

Ph.
C
Ph.
B DC

Ph.
A

Current Coolant
sensors in/out

(a) dimensions in mm (b)

Fig. 7.4: Prototype SiC inverter with XM3 half-bridge power modules: (a) CAD drawing
showing its outer dimensions and (b) the final assembled inverter

Considering the above dimensions of the inverter and a peak power of 250 kW, the
resultant power density is 250kW/5.107liters = 48.9 kW/liter. The weight of the
inverter is estimated to 5.1 kg through calculations on the CAD model, therefore
the power-to-weight ratio is 49 kW/kg. These numbers are within the usual power
densities reported in [177] for SiC inverters. It should be noted that the MCU
board for controlling the switches and the housing of the inverter are not included
in these calculations.
The switching transient of the SiC MOSFETs is examined in Fig. 7.5 in order to
evaluate the inverter design. The drain-source voltage VDS of the bottom switch of
phase B is shown when 300 V dc-link voltage is applied and 148 A current passes
through. As seen in the Fig 7.5(a), there is a low voltage overshoot of 25.9 V
(8.6% of the applied dc voltage) and dv/dt equal to 7.4 kV/µs during the turn-off
switching transient. The low voltage overshoot verifies that the designed dc-link
has successfully maintained low parasitic inductance between the capacitors and
the switches. The dv/dt has been calculated with the following formula, as cited
from [180]
dv 0.8VDS peak
= (7.9)
dt tr
where tr is defined as the time required for the voltage to rise from 10% to 90% of

152
7.4. Design of Liquid Cooling System for Inverters

X 25.36
Y 325.9 X 33.59
X 25.35
Y 304.2 X 33.71
Y 293.8
300 300 Y 272.8

VDS [V]
VDS [V]

200 200

100 100
X 33.76
X 25.15 Y 21.68
Y 1.601 X 25.31
0 Y 33.93 0
X 33.77
Y -9.965
25 25.2 25.4 25.6 25.8 32 32.5 33 33.5 34 34.5 35

(a) (b)

Fig. 7.5: Experimental measurements of the drain-source voltage of phase B bottom


switch with 300 V dc voltage and 148 A current during (a) turn-off and (b)
turn-on transient

its peak value. The measured rise time equals 35.1 ns for the turn-off transient
shown in Fig. 7.5(a). External gate resistance RG = 1 Ω (additionally to the
internal gate resistance of the module 2.5 Ω) is used for this inverter.

7.4 Design of Liquid Cooling System for Inverters


Optimizing the heatsink design of a liquid-cooled automotive inverter is an impor-
tant step that can lead to more compact drivetrains with least possible material
expense and longer life cycles. This would allow the development of more efficient
packaging of automotive drives [181, 182].
The standard package layout of wire-bonded power modules used nowadays in
high-power inverters is composed of multiple layers of materials between the die,
which needs to be cooled, and the heatsink, in order to electrically isolate these
two points. This structure is called direct bond copper (DBC) and consists of the
following layers [183, 184]:
• a copper layer is soldered beneath the the semiconductor die,
• the ceramic substrate is placed beneath, which is made of an electrically
insulating and thermally conductive material, such as the AlN,
• a second copper layer follows and
• the baseplate is soldered at the bottom of the whole module, which consists
of a thicker piece of copper or aluminium. The heatsink is placed at the
bottom side of the baseplate, which is called the case of the module
There are power modules recently released that do not have a baseplate, but

153
Chapter 7. Electrical and Thermal Design of Automotive SiC Inverter

the heatsink is connected directly to the bottom side of the second copper layer,
such as the Easy power module family from Infineon [185] and the WolfPACK
from Wolfspeed [186]. The main advantages of this structure is the reduced cost
of the device due to the absence of the thick baseplate and the better thermal
conductivity [187].
In most of the available literature, heatsinks are dimensioned for maximum load,
as in [188]. However, the inverter may never reach these limits under normal
driving conditions of a passenger BEV. Despite this, steady-state CFD simulations
are often utilized in similar thermal modelling studies [181, 189–194], which do not
represent the reality, since the load of the drive is constantly varying.
Long real-time thermal simulations of the inverter heatsink are necessary in order
to evaluate the temperatures of the designed system under transient load and to
evaluate its efficiency and reliability. These simulations are implemented with
equivalent and simplified models in [195–198], which cannot represent the thermal
behavior of the semiconductor devices and the coolant flow inside the heatsink.
Quasi-transient CFD modelling has also been used in [199, 200] in order to reduce
the computational load. However, the results are not as realistic as in the fully
transient 3D CFD approach.
The sizing of the heatsink involves a multidisciplinary approach in order to capture
all the physical properties for thermal conditioning of the electric components.
The thermal performance of the inverter can be studied more accurately with
3D Conjugate Heat Transfer (CHT) computations, where both fluid and solid
elements in thermal contact are handled. The turbulence of the fluid, which is a
strictly 3D mechanism, should also be modelled for accurate results. From the
above it is concluded that an effective and practical thermal design technique,
specially tuned for the cooling component of automotive SiC high-power inverters,
is missing from the technical literature. Time-varying 3D CHT modelling has also
not been utilized yet as the design tool for the inverter heatsink.
This section presents a new iterative heatsink design tool based on the transient
3D CHT computation technique, where the cooling plate geometry of the SiC
inverter in Fig. 7.4(a) is fine-tuned in order to meet the specific needs of a BEV.

7.4.1 Overview of Cooling Plate Geometries


The main objectives for the heatsink design are
• to have good heat dissipation ability

154
7.4. Design of Liquid Cooling System for Inverters

• the pressure drop of the coolant to be kept as low as possible, since the
inverter is usually cooled in the same cooling loop with the motor. A typical
limitation for an automotive inverter can be set to 150 mbar, as referenced
from [8].
• to have uniform heat dissipation at the contact surface with the power
modules, so that the case-surfaces of all the three power modules experience
similar thermal conditions and, therefore, the same aging. Inhomogenous
cooling of the semiconductor chips would lead to a temperature gradient
[191, 201] affecting the lifetime of each switch [202].
• manufacturability with low-cost techniques [192], such as machining and
forging, without using special tools or more expensive solutions.
Cooling plates with simple straight channels extruded or machined in the aluminium
plate have been traditionally been used in commercial inverter heatsinks [191, 194].
These designs have simple structure and, consequently, low manufacturing cost.
However, they cannot be easily optimized for high power density applications
where advanced thermal characteristics with low pressure drop of the coolant are
required.
Therefore, heatsinks with more sophisticated structure, accommodating multiple
fins, have been developed in order to enhance the heat transfer capability between
the semiconductors and the coolant. These fin geometries can be found in the
design of the baseplate for direct-cooled power modules [203], as well as in cooling
plates for power module packages with flat baseplate that require thermal grease
for better heat conductivity. There are heatsinks with straight [193] and wavy
[204] fin shapes that manage to enhance the heat transfer coefficient between the
plate and the coolant, due to increased mixing of the coolant liquid.
Plate designs with pins of different shapes and sizes have also shown to attain good
cooling performance with high convective heat transfer coefficient. Pin-fins with
round shape are studied in [205] and rectangular pins are shown in [191]. More
advanced pin-shapes are proposed by Danfoss [191] and MicroCool [192], which
also require a more special, custom manufacturing process. A certain method to
optimize the shape of the fins and inner layout of the cooling plate needs to be
followed for each study, using a parametric analysis of the key dimensions of the
design [201] or iterative optimization algorithms as in [181].

155
Chapter 7. Electrical and Thermal Design of Automotive SiC Inverter

7.4.2 CHT Computation Approach with Steady-State Heat


Load 2
At this step of the design process, multiple heatsink geometries are investigated
and simulated with constant heat dissipation. A total power loss of 4 kW has been
selected for all the three power modules and coolant flow speeds vary between
2.5 l/min and 10 l/min. A full 3D CFD model is used to simulate the heatsink of
the inverter. The simulation toolbox is OpenFOAM [206] and the solver is named
chtMultiRegionSimpleFoam, which is suitable for steady-state CHT computations.
The investigated computational domain comprises individual meshes for the coolant
and the heatsink, as shown in Fig. 7.6(a). The following input parameters have
been considered:
• The coolant is a mixture of 50% water and 50% ethylene glycol and the solid
material is aluminium.
• The coolant inlet temperature is selected equal to 300 K (or 26.85o ) for the
results of this subsection.
• The developed inverter is intended to be used together with the liquid-cooled
PMSM studied previously in Chapter 6, which has coolant inlet nozzle with
inner / outer diameter 14 / 19 mm.
• The outer dimensions of the inverter heatsink are L × W × H = 193.5 × 85 ×
37 mm. Further parameters, such as mass and material properties of the
heatsink, are provided by the CAD drawing of the heatsink [see Fig. 7.4(a)].
The boundary conditions of the simulation are designed so as to represent the
three power modules of the three-phase inverter as surface patches on the sink that
have an evenly distributed heat source, with total thermal power of each patch
equal to 4/3 kW. The mean temperature of each of these rectangular patches,
being averaged at the surface of the heatsink, is shown as Tsw1 up to Tsw3 in the
following simulation results.
The first design candidates is a standard cooling plate with straight cooling
channels as shown in Fig. 7.7. In order to be able to fit 2-loops of the straight
rectangular channels in the given cooling plate dimensions, the inlet diameter has
been selected to be equal to d1 = 9 mm and the spacing between the channels
2 More details can be found:

G. Mademlis, R. Orbay, Y. Liu, and N. Sharma, "Designing Thermally Uniform Heatsink with
Rectangular Pins for High-Power Automotive SiC Inverters," in Proc. IECON 2020 - 45th
Annual Conference of the IEEE Industrial Electronics Society, 2020, pp. 1317-1322.

156
7.4. Design of Liquid Cooling System for Inverters

HEAT SOURCE
POWER MODULE 3
POWER MODULE 2 HEAT SOURCE
POWER MODULE 1 HEAT SOURCE

(a)
(b)

Fig. 7.6: First investigated heatsink candidate showing (a) the cooling plate (grey)
and the cooling circuit (blue) and (b) the boundary conditions for the CHT
modelling of the heatsink.

is b1 = 5.42 mm. A straightforward disadvantage of this heatsink design is the


different nozzle diameter compared to the motor of the drive. As seen also in Fig.
7.7(b), there is a temperature gradient on the surface of the heatsink, with the
power module 1 on the right side experiencing the highest temperature.
The second heatsink geometry studied is with multiple fins placed along the flow of
the coolant similar to the design presented in [207] and it is shown in Figs. 7.8-7.9.
The inlet diameter for all the following designs is 19 mm and the coolant enters
from the left side of the cooling plate and exits at the right side. The design in Fig.
7.8 has four wave-shaped fins in order to increase the surface area between the
coolant and the plate and to introduce turbulence which would further enhance
the cooling performance. The main challenge in this design is to split the coolant
flow after it enters the plate into the five resultant channels. This is not done

(a) (b)

Fig. 7.7: Initial heatsink design candidate with a 2-loop cooling channel and coolant flow
10 l/min.

157
Chapter 7. Electrical and Thermal Design of Automotive SiC Inverter

adequately in the first design of Fig. 7.8, since most of the coolant keeps flowing
in the middle channel.

(a) (b)

Fig. 7.8: Heatsink design "wave 1", with multiple fins of wavy shape.

A modified version after some more design iterations is shown in Fig. 7.9, that
achieves a more even flow of the coolant through most of the channels. This
improvement is also reflected on the sink temperatures Tsw shown in Fig. 7.9(b),
which are considerably lower compared to the temperatures of Fig. 7.8(b). However,
the same temperature gradient with the initial 2-loop design is evident here with
the Tsw1 being the highest among the three.

(a) (b)

Fig. 7.9: Heatsink design "wave 2", having multiple fins of straight and wavy shape.

Heatsink geometries with pins have also been studied and an example is presented
in Fig. 7.10. Rectangular pins are used due to their simpler and cheaper design,
since they can be machined [191], and their better thermal performance [208].
The number of pins, their positions and other parameters are optimized through
iterative simulations. The presented heatsink design has a more homogenous
temperature distribution per MOSFET patch compared to the other candidates,
as seen in Fig. 7.11(a), and has quite good thermal performance and pressure
drop, as per Figs. 7.11(b)-(c). Therefore, the heatsink with rectangular pins has
been selected for further analysis in the next subsection.

158
7.4. Design of Liquid Cooling System for Inverters

Module 3 Module 2 Module 1

Tsw

(a) (b)

Fig. 7.10: Heatsink designs with multiple rectangular pins and three straight fins dividing
the flow into equal sections
Module temperature Tsw [°C]
(Tsw )

20 120 0.2
Tsw1 "2-loop" design
Tsw2 "wave 1" design
10 110 0.15 rectangular pins

P [bar]
Tsw3
Standard deviation

5 100
rectangular pins 0.1
"wave 1" design
"wave 2" design 90
0.05
80
1 0
4 6 8 10 4 6 8 10 4 6 8 10
Flow speed [l/min] Flow speed [l/min] Flow speed [l/min]
(a) (b) for the heatsink with pins (c)

Fig. 7.11: Comparative simulation results of the studied heatsinks: (a) standard deviation
in logarithmic-scale showing the discrepancy between the three power module
temperatures; (b) average temperature of the power module patches for the
heatsink with pins; (c) pressure drop of the coolant (between inlet/outlet)

7.4.3 CHT Computation Process with Transient Heat Dis-


sipation
The results of the final heatsink design with rectangular pins (Fig. 7.10) are shown
in this subsection, while the load of the modelled inverter is dynamically changing.
The performance of the designed heatsink is evaluated, when typical operating
conditions for automotive drives are simulated. Firstly, the acceleration test of a
test-case BEV (with parameters displayed in Table 7.2) is performed, when the
peak torque of the drive equal to 580 Nm is applied. Then, simulation results of
the motor drive with the CLTC_P driving cycle are presented.
The test-case vehicle is simulated first on Matlab/PLECS, where detailed repre-
sentation of the vehicle dynamics as described by (7.1)-(7.5), the motor and the
SiC inverter model are included. As seen in Fig. 7.12, except for the rotor inertia
of the PMSM, the inertia due to the mass of the vehicle m is considered, which

159
Chapter 7. Electrical and Thermal Design of Automotive SiC Inverter

converted to the motor side is calculated as


m r2
Jcar = (7.10)
gear2

The purpose of the Matlab/PLECS model is the calculation of the inverter losses,
which is the heat-load of the studied heatsink. An equivalent thermal network of
the inverter is also developed in order to calculate the losses of the power switches
at the correct temperature that the inverter experiences. The results from the
steady-state CHT computations of the previous subsection can be utilized to build
an equivalent thermal model of the inverter’s cooling system, which is a combined
Cauer and Foster network, cited from [202].
The variable Rth sa in Fig. 7.12 represents the thermal resistance between the
aluminium cooling plate and the coolant fluid. Rth cs is the thermal resistance
between the case and the sink, because of the thermal interface material (TIM). The

Table 7.2: Modelled Passenger Vehicle Parameters


PWM
Parameter Value Unit Foster Network Vdc
PMSM
Vehicle mass m 1900 kg Junction
R
Case
Tc [°C]
800V

Tj th jc -
Effective area A [162] 2.2879 m 2
A [°C]
+

Cth j
Aerodynamic drag Cd [162] 0.35 - Jcar loss
T
Periodic Rth cs Rth sa Ta [°C]
c

Rolling resistance Cr [162] 0.012 - Tload [Nm] average + -


Cth s
Tire size 245/45R19 Vehicle model: P loss
[W]
Wheel radius r 0.352 m (6.1)-(6.5)
θe [ ] o
Cauer Network

Powertrain arrangement 1 motor idq [A] Vector


*
vdq
PWM
Control
Peak torque Te max 580 Nm
*
* i dq Voltage
Te MTPA limiter
Gear ratio gear 5.7 -
[Nm] Current
limiter

Maximum speed vmax 209 km/h


Fig. 7.12: Schematic of the modelled electric
Number of pole pairs p 4 - drive on Matlab/PLECS
Stator resistance Rs 0.026 Ω
Stator inductance Ld 0.90 mH
Stator inductance Lq 2.61 mH
DC-link voltage Vdc 800 V
PMSM shaft inertia J 0.018 kg·m2
Inverter switching freq. fsw 20 kHz
Inverter gate resistance RG 1 Ω
Case-sink Rth cs sw [209] 0.026 K/W
Heatsink mass mh 1.36 kg

160
7.4. Design of Liquid Cooling System for Inverters

TIM affects the heat spreading in the module and, therefore, Rth cs is important
to be included in the thermal simulation model. Rth jc is the thermal resistance
between the junction of the power switch and the case, because of the different
layers of the DBC structure of the module. The Rth jc has been calculated by a
thermal model provided by the switch manufacturer.
L Rth cs sw [209]
Rth cs = = = 4.3 · 10−3 K/W (7.11)
k Acontact nsw nsw =6

1
Rth sa = (7.12)
hsa Ac
where Acontact is the total area of the power modules’ baseplate, L the thickness
of the TIM. The Rth cs value is not available for the specific power modules and
instead its value is taken from [209] for IGBT power modules of similar dimensions
and current rating. Rth cs sw = 0.026 K/W is the cited case-sink resistance per
individual switch position of the inverter. Ac = 0.0604 m2 is the total surface
of the heatsink cooling channel in contact with the coolant and hsa is the heat
transfer coefficient of the heatsink with rectangular pins, which is obtained from
the CHT results of the previous subsection.
The thermal capacitance of the cooling plate is calculated as
cp Al
Cth s = cp mh ==⇒ Cth s = 1238 J/K (7.13)

where cp Al is the specific heat capacity of aluminium equal to 910 J/(K kg).
The conduction losses of the SiC MOSFETs and their body diodes are calculated
with the PLECS block periodic average, while periodic impulse average is used for
the switching losses. As seen in the schematic of Fig. 7.12, the equivalent thermal
model is a closed loop system, due to the feedback of the inverter’s temperature
in the power loss calculation and, therefore, the losses are accurately estimated at
the correct temperature for each simulated operating point.

Acceleration of the BEV with maximum torque:


The acceleration profiles under-test are analysed for inlet coolant temperatures
Tin equal to 26.85°C (300 K) and 65°C, in order to model driving conditions,
when the BEV operates at a cool or a warmer environment, respectively. It
has been assumed that the Tin remains constant throughout the whole test
procedure. Conforming with the usual practice in light duty vehicle cooling
systems, 6.8 l/min coolant volume flow is used, similar to [210]. The suitable heat
transfer coefficients hsa are obtained from the steady-state CHT results and are
equal to hsa 26.85o C = 2228.41 W/(m2 K) and hsa 65o C = 2279.21 W/(m2 K).

161
Chapter 7. Electrical and Thermal Design of Automotive SiC Inverter

The motor accelerates at the beginning and constant torque is applied equal to
Tm max = 580 Nm. When the current controller reaches its voltage limit and
enters the medium- and high-speed operating regions, the applied torque starts to
reduce. Then, the motor operates with constant voltage equal to its maximum
value, while the current changes. At t = 5.2 s the system reaches its maximum
power and the inverter experiences also its maximum losses and temperature.
The results from the acceleration test are shown in Fig. 7.13 as obtained from
Matlab/PLECS, showing that the modelled BEV can accelerate to 100 km/h in
6 s. The semiconductor losses are shown in Fig. 7.13(c) for both tested coolant
temperatures and it can be observed that the increase of the temperature Tin
leads to higher losses of up to 16.3%.
Car accelaration [m/s2]

6
PMSM torque Te [Nm] 600
200 X 21.8
Car speed [km/h]

X 11.2
Y 149.9 Y 200
150 4 500
X6
Y 100.2
100 2 400

50 0 300

0 -2 200
0 5 10 15 20 0 5 10 15 20
Time [s]
(a) (b)
X 5.2
8000 250 Y 231.8
with Tin= 65 oC
Power losses [W]

X 5.2
Y 6198 X 4.7
6000 200 Y 179.6
with Tin= 26.85oC
Power

150 Active Power Pe [kW]


4000 Reactive Power Qe [kVAr]
100
2000
50
0 0
0 5 10 15 20 0 5 10 15 20

(c) (d)

Fig. 7.13: Results of the (a) speed and acceleration of the BEV, (b) applied electromag-
netic torque, (c) semiconductor losses of the inverter for different inlet coolant
temperatures Tin and (d) active / reactive input power of the motor

Then, the inverter is simulated with the 3D CHT model on OpenFOAM and
input of the model is the power-loss-curve from Fig 7.13(c). The results from the
transient 3D CHT simulations are shown in Fig. 7.14 for the same acceleration
profile. The hottest temperature of the heatsink is reached at 10 s. The junction
temperatures for each individual power module Tj0 are also shown in Fig. 7.14
with dotted curves. The discrepancy between the individual junction temperatures
of the three power modules can be observed. For example, as can be seen in Fig.
7.14(a), at 10 s the temperature rise of Tj2
0
(with respect to the Tin ) is 7.4% and
8.8% higher than the temperature rise of Tj1 0
and Tj3
0
, respectively. In a heatsink

162
7.4. Design of Liquid Cooling System for Inverters

without uniform heat distribution, such as the one in Fig. 7.7, this discrepancy
could also be a potential cause of failure for the inverter and faster aging of the
hotter semiconductors.

΄
΄

΄
΄
΄

2s 3s 4s 2s 3s 4s

5s 10 s 15 s 5s 10 s 15 s

Tsw [oC] : Tsw [oC] :

(a) (b)

Fig. 7.14: 3D CHT results during acceleration of the BEV with maximum motor torque,
when (a) Tin = 26.85°C and (b) Tin = 65°C

The heatsink temperatures in Fig. 7.14, which are simulated with the proposed
transient method, can be compared with the ones obtained by the steady-state
technique (straight dashed-dotted lines), which have been shown previously in Fig.
7.11(b) for 6.8 l/min coolant flow speed. As can be seen, the sink temperature
values calculated with the steady-state simulations are higher than the ones
obtained with the transient CHT computations, throughout the whole acceleration
phase of the vehicle. It should be noted that the steady-state calculations refer
to 4 kW semiconductor losses, while the computation results by the proposed
transient 3D CHT design technique correspond to dynamic power loss of the SiC
inverter up to 6.2 kW [Fig. 7.13(c)]. Designing the heatsink with steady-state CHT
simulations to have 6.2 kW of thermal load would lead to an oversized cooling
plate, whereas the design with the new transient technique has been accurately
tailored to the exact acceleration requirements of the examined electric vehicle.
Therefore, it gives the margin for further dynamic overload with additional losses
even up to 55%. This means that designing the heatsink with transient CHT
simulations allows better exploitation of the inverter at dynamic operation.

Driving Cycle Simulations with Transient CHT:


Simulation results of the tested PMSM drive from the CLTC_P driving cycle are
shown in Fig. 7.15. The inlet coolant temperature in both cases is set equal to
Tin = 26.85°C and the coolant volume flow is 6.8 l/min. The semiconductor losses

163
Chapter 7. Electrical and Thermal Design of Automotive SiC Inverter

and temperatures are shown in these figures.


Fig. 7.15(a)-(b) shows the inverter power losses and temperature, when the electric
drive is simulated on Matlab/PLECS. Similarly with the acceleration test, these
power losses are imported into the 3D CHT model as variable heat-load of the
heatsink and the system is simulated again. A closer look of the first 271 s of
the CLTC_P cycle is shown in Fig. 7.15(c) showing the average temperature
of each power module patch Tsw and screenshots of the cooling plate for some
characteristic time moments. The junction temperatures Tj0 are shown in Fig.
7.15(d).

(a) (b)

45
41
40
39
38
37
36

Tsw [oC]
35
34
33
32
31
30
29
28
27

20 s 30 s 75 s

90 s 220 s 250 s
(c)
60 =52.39oC
Junction temp. [oC]

j1
Equiv. model (tuned):
j1

j2
=54.05oC Equiv. model (tuned): j2
50 =51.74oC Equiv. model (tuned): j3
j3

40

30

0 50 100 150 200 250


Time [s]
(d)

Fig. 7.15: CLTC_P driving cycle: results of the (a) total inverter losses and (b) heatsink
temperatures from the thermoelectric system model and (c) individual power
module temperatures obtained from the 3D CHT analysis with Tin = 26.85°C
and (d) junction temperatures of the three power modules

164
7.4. Design of Liquid Cooling System for Inverters

The temperature Tsw2 is the highest most of the time, except for some time
instants during the deceleration phase of the cycle such as at 50 s and at 90 s.
Then, the Tsw2 is the lowest temperature among the three. The same results can
also be seen for the junction temperatures. The highest temperature discrepancy
between the three power modules is seen at 20 s, when the temperature rise of Tj20

(with respect to the Tin ) is 6.5% and 9.3% higher than the temperature rise of Tj1
0

and Tj3 , respectively.


0

Another interesting observation is that the hottest area of the cooling plate
(pinpointed with a purple dot) is close to the middle of the device, while the
temperature keeps increasing until it reaches its peak value, such as at the time
moment 20 s when the hotspot-temperature is 43o C. Afterwards, when the temper-
ature starts to decrease, the hottest area moves towards the bottom of the plate,
such as at 30 s when the purple dot is 20 mm further down. Similar observation
can be made for the acceleration speed profile in Fig. 7.14. The consequence
of this is that, depending on the operation of the vehicle different areas of the
power modules experience the highest temperature affecting also the lifetime of the
corresponding semiconductors. Therefore, the transient 3D CHT computations
are important in order to study these thermal heterogeneities and to try to tune
accordingly the inverter design.

7.4.4 Iterative Cooling Design Process


The following thermal design proceedure for automotive inverters can be formulated
as a conclusion from the simulation results of this section. Firstly, the needed ther-
mal dissipation power of the inverter’s heatsink is defined with a Matlab/PLECS
analysis for constant load of the drive. Drive operation close to its maximum power
can serve as a starting point for the iterative design process calculating the losses
at the worst-case conditions for the inverter. The initial thermal properties of the
cooling plate are decided with a hand-computed Nusselt-number-based approach
[211]. The required thermal dissipation power Ploss init is calculated during this
design step, where the first draft of the heatsink design will be based on. The
Ploss init is being updated during later steps of the iterative design process and a
more accurate value is later defined, as shown in Fig. 7.16.
Afterwards, a Computer Aided Design (CAD) of the heatsink is prepared for
further computations. A full 3D CHT model of the heatsink is then used in the
next step using the software OpenFOAM, where the performance of the design
is evaluated for various operating conditions of the fluid, such as different flow
speeds and cooling media temperatures. In case the designed heatsink fulfills the

165
Chapter 7. Electrical and Thermal Design of Automotive SiC Inverter

Equivalent thermal model:


maximum constant load
Matlab / PLECS

Ploss init
estimation

CAD of the
cooling plate

heatsink design
.stl format modify
3D CHT for design
Subsection 6.4.2 constant Ploss OpenFOAM

T within
Yes No
constraints
2
hsa [W/m K]

Matlab / PLECS

Subsection 6.4.3 Equivalent


Equivalent
thermal model:
thermal model: Other tests with
Transient load
tuning Transient load transient load
vehicle
thermal driving cycle
acceleration
model
Tsw 1,2,3(t)
Tmax , ΔΤ and ΔP within
Yes No
constraints
Ploss(t)

3D CHT for final design


Ploss(t) reached

OpenFOAM

Fig. 7.16: Flowchart of the inverter’s heatsink design process adopting an equivalent
thermal and electric model and 3D CHT analysis.

requirements of the specific application (e.g. the temperature of the inverter lays
withing the predefined safety limits), the design process moves to the next step
which includes testing of the inverter with transient load.
The results of the steady-state 3D analysis are utilized to build a simplified thermal
network model of the inverter, where the power losses at various operating points
can be calculated. Many driving patterns of the BEV can be tested here with the
modelled motor drive, such as an acceleration of the car from standstill up to its
maximum speed or legislative driving cycles. If the calculated temperatures of the
power modules during these tests lay within the recommended operating conditions
defined by the MOSFET manufacturer, the power loss profile is used by the 3D
model of the heatsink to run transient CHT computations with the same load

166
7.5. Inverter Efficiency Measurements

patterns, where the individual temperature of each power module can be obtained.
In case the calculated temperatures or the pressure drop of the heatsink’s coolant
lay outside the design specifications, the design process is repeated, updating
accordingly the value of Ploss init in the first step of the flowchart. The heatsink
design is refined by this algorithm with the constant and transient load tests until
all the design specifications are fulfilled.

7.5 Inverter Efficiency Measurements


The efficiency of the SiC inverter has been measured experimentally with switching
frequency fsw = 10 kHz and 20 kHz and the results are shown in this section. The
60 kW PMSM emulator set-up of that has been studied in Chapter 6 is used for
the measurements [see the experimental set-up in Fig. 6.8(b)]. The inverter is
cooled with cold tap water and an automatized process is used for measuring the
voltages and currents of the system at each operating point, so that the inverter is
loaded for a short time and and constant temperature can be kept throughout the
whole experiment. A dead-time equal to 1 µs has been used between the top and
bottom switch of each phase leg of the inverter.

P 13.49
100 n10 98.6

P 20.62
n20 98.1
Efficiency [%]

95

90

85 fsw=10 kHz
fsw=20 kHz
80
0 10 20 30 40 50 60
Inverter output power [kW]
(a) for fsw = 20 kHz (b)

(c) for fsw = 10 kHz (d) for fsw = 20 kHz

Fig. 7.17: Experimental measurements of the inverter (a) output power, (b) efficiency at
3000 rpm, efficiency map (c) for 10 kHz and (d) 20 kHz switching frequency
with dc-link voltage 360 V

167
Chapter 7. Electrical and Thermal Design of Automotive SiC Inverter

The operating points on the torque-speed map shown previously in Fig. 6.10 are
tested with the PMSM emulator. However, due to the voltage drop across the ac
inductors of the topology, the points where the machine operates with maximum
stator voltage at the high-speed region cannot be emulated. Also, a slightly larger
dc voltage than the one examined in Chapter 6 has been used here, namely 360 V
instead of 300 V.
The active power at the ac side of the inverter, which has been tested here, is shown
in Fig. 7.17(a). The envelope of the torque-speed map is also shown with black
line in the same figure. The inverter efficiency in Figs. 7.17(b)-(d) is calculated as
follows
Pac
Efficiency% = 100 (7.14)
Pac + Pinv loss
where Pac is the inverter active power at its ac side and Pinv loss is the power loss
of the inverter as measured with the oscilloscope.
The inverter efficiency has a peak value of 98.6% for switching frequency 10 kHz
and 98.1% for switching frequency 20 kHz, as seen in Fig. 7.17(b). The largest
difference between these two graphs can be observed when low power is tested,
while similar inverter efficiency for both cases can be observed when the power
increases. The measured efficiency complies with previously reported results in
the literature for SiC three-phase inverters [212, 213].

7.6 Summary
The design process of an automotive inverter has been presented in this chapter.
SiC power switches were used in the inverter in order to reduce its losses and
increase its power density. Prototypes of two inverters were manufactured that
were used in the PMSM drive which has been tested experimentally in Chapter 6.
The first step involves the formulation of the design requirements for the inverter,
as imposed by the expected capabilities of the vehicle where the inverter is intended
to be used. The modelling process of the vehicle dynamics is described, which
determines the required power of the motor drive.
The next step is the dimensioning of the inverter and the selection of its components,
based on the design specifications. The power switches and the dc-link capacitors
are the basic components that should be selected.
The last step is the design of the cooling system of the inverter and, specifically, its
heatsink. A multidisciplinary design process of the heatsink is proposed here which

168
7.6. Summary

consists of multiple simulation stages. The inverter losses are modelled through
circuit simulation tools, such as the PLECS toolbox on Matlab. Then, the heatsink
is designed through CAD and simulated through CHT computations, firstly with
steady-state heat-load and later with transient load considering realistic operating
profiles of the vehicle. The main advantages of the transient CHT modelling
approach can be summarized as follows:
1. Firstly, the overload capacity of the inverter can be more easily evaluated
with the proposed design method, when the maximum acceleration of the
investigated vehicle is applied. The results of the simulated test-case system
have shown that short-time overloading with an increase of the thermal heat
up to 55% still keeps the inverter heatsink temperature beneath the safety
limits.
2. The transient CHT model can reveal more accurately the temperature
discrepancies among the three SiC power modules during realistic operating
conditions, which can reach up to 9.3% for the tested automotive inverter.
3. The location of the hotspots at the surface of the sink can be more accurately
calculated with the transient 3D design method. This can allow more fine-
tuned design of the heatsink in order to reduce large temperature differences
at the surface of the sink.
This procedure has as a result a fine-tuned heatsink design that fulfills at the same
time the design specifications of the specific vehicle with the smallest possible
component size.
Lastly, experimental measurements of the prototype inverter efficiency are pre-
sented for different power and frequency operating points. When the inverter
operates with switching frequency 20 kHz and 360 V dc-link voltage, the maximum
efficiency is 98.1%.

169
Chapter 7. Electrical and Thermal Design of Automotive SiC Inverter

170
Chapter 8

Conclusions and Future Work

8.1 Conclusions
The design and control of SiC three-phase inverters for machine drives have
been investigated in this thesis. Two background applications have been studied,
kite-based tidal power generation and battery electric vehicles. Although these
applications are very different from each other, they share similar goals and
challenges for the design of the power conversion system, namely high power
density and efficiency.
Two different approaches have been investigated in order to achieve high power
density of the power conversion system. Firstly, the increase of the operating
voltage of the machine drive, which is used in offshore tidal power systems, has
been motivated by the resultant decrease of the ac current flowing through the
undersea cables that connect the kite to the on-shore grid. Since the use of medium
voltage drives with 5-level converters can increase the operating voltage of the
system more than four times compared to standard low-voltage 2LCs, undersea
cables with smaller diameter can be used, reducing in this way the total weight
and drag of the kite system. Therefore, a medium voltage multilevel generator
drive has been proposed in Chapter 4, which consists of two B2B-connected 5-level
NPC converters and a medium voltage PMSG. The 5-level NPC converter has
been proven to be a suitable topology for medium voltage grid-connected machine
drives, because its main advantages compared to other multilevel topologies are
the common dc bus, which is convenient for B2B configurations, and the reduced
number of passive components. Conventional vector control can also be used to

171
Chapter 8. Conclusions and Future Work

control both the machine inverter and the grid inverter.


The harmonics of the ac current and voltage are also highly improved when using
drives with 5-level converters instead of 2LCs. This constitutes an additional
motivation towards the use of multilevel converters in offshore power plants, since
the size of the ac passive filters can be drastically reduced and higher power
density of the drive can be attained. Specifically, FFT analysis of the current and
voltage has shown that the THD of the generator current is reduced by 66% and
the generator phase voltage by 58%, when 5-level instead of 3-level converters
are used. Experimental results in Chapter 5 have shown similar results. Similar
improvement is observed also on the grid-side voltage and current harmonics.
The main drawback of this converter topology is the balancing of the dc-link
capacitor voltages. This problem has been solved in the thesis by proposing four
alternative ways of balancing the dc-link voltages, which utilize a modified version
of the SVM or dc/dc converters connected in parallel to the dc-link. The advantages
and disadvantages of each proposed balancing method have been summarized in
Table 4.8 and the conclusions mentioned there are based on simulation results
implemented in Matlab/Simulink and PLECS.
The second approach towards high power density involves the use of SiC power
switches in the converters. SiC switches have lower losses compared to standard Si
IGBTs, which allows to build more compact inverters and to use higher switching
frequencies.
The mechanical design of the SiC inverter is an important step of the whole
manufacturing process. An optimization procedure of the cooling components
using transient CFD computations and the correct sizing of the dc-link capacitor
bank have been introduced and applied on a test-case 450 A, 800 V (dc) SiC
inverter. The proposed cooling design approach allows an evaluation of the overload
capability of the inverter, which can reach up to 55% for the test-case inverter,
and a more accurate estimation of the temperature distribution on the inverter
heatsink.
A prototype 450 A SiC 2-level inverter has been manufactured and tested exper-
imentally on a 60 kW PMSM test-bench. Afterwards, a P-HIL set-up has been
studied, where the 60 kW PMSM has been emulated using an additional converter.
P-HIL testing for machine drives is shown to be a useful experimental verification
method for inverter design projects, in case multiple machine types need to be
evaluated with the same inverter and the machines are not available yet.
However, circulating zero-sequence current is a typical issue of the examined

172
8.2. Future Work

machine emulator topology, which is usually solved by installing hardware common


mode filters. A novel current control scheme has been proposed that can limit
the circulating current that flows in the system by controlling the duty cycle
distribution among the zero states on the SVM. The effectiveness of the control has
been verified experimentally, since the remaining zero-sequence current amplitude
has been successfully limited to 3.3% of the ac current amplitude, when the system
operates with 42 kW active power at the base speed of the modelled PMSM.

8.2 Future Work


The following research points can be further investigated in future projects in
order to extend and further utilize the outcomes of this PhD thesis:
• The use of multilevel-based drives has been investigated only for offshore re-
newable power generation in this thesis. It would be interesting to investigate
the possible advantages of using 3- or even 5-level inverters in automotive
drives. Especially, future electrification of heavy-duty or construction ve-
hicles with large power demands would be benefited by an increase of the
powertrain voltage and multilevel inverters would be a viable solution to
realise these systems.
• The proposed zero-sequence current control for P-HIL systems has been
applied only on 2LCs, where two redundant zero states exist. However, it is
possible to extend the validity of this control to the multilevel space vector
plane, where more zero-state combinations exist. This would be necessary
in case multilevel inverters are tested in P-HIL systems.
• Integrated machine-inverter systems, where the power switches of the inverter
are mounted directly on the housing of the machine, are considered as
a promising solution for future extremely high-power-dense drives. The
inverter design approach proposed in this thesis, which uses transient CFD
computations to model the cooling system, could be applied in such a project
improving the thermal properties of the resultant drive.

173
Chapter 8. Conclusions and Future Work

174
References

[1] A. Khaligh and M. D’Antonio, “Global trends in high-power on-board charg-


ers for electric vehicles,” IEEE Trans. Vehicular Technology, vol. 68, no. 4,
pp. 3306–3324, 2019.
[2] IEC 61851-23:2014, “Electric vehicle conductive charging system - Part 23:
DC electric vehicle charging station,” Standard, March 2014.
[3] IEC 62196-3:2014, “Plugs, socket-outlets, vehicle connectors and vehicle
inlets - Conductive charging of electric vehicles - Part 3: Dimensional
compatibility and interchangeability requirements for d.c. and a.c./d.c. pin
and contact-tube vehicle couplers,” Standard, June 2014.
[4] SAE J1772, “Electric Vehicle and Plug in Hybrid Electric Vehicle Conductive
Charge Coupler,” Standard, January 2010.
[5] D. Pehrman, “Design Aspects of Inductive Power Transfer Systems for Elec-
tric Vehicle Charging,” Licentiate Thesis, Chalmers University of Technology,
2019.
[6] N. Keshmiri, D. Wang, B. Agrawal, R. Hou, and A. Emadi, “Current status
and future trends of GaN HEMTs in electrified transportation,” IEEE Access,
vol. 8, pp. 70 553–70 571, 2020.
[7] M. Ehsani, Y. Gao, and J. M. Miller, “Hybrid electric vehicles: Architecture
and motor drives,” Proceedings of the IEEE, vol. 95, no. 4, pp. 719–728,
2007.
[8] L. Tribioli, “D5.1 Documentation of the basic concept of the cooling circuit,”
Drivemode - Horizon 2020, Project report, 2019.
[9] X. Kuang, K. Li, Y. Xie, C. Wu, P. Wang, X. Wang, and C. Fu, “Research
on Control Strategy for a Battery Thermal Management System for Electric
Vehicles Based on Secondary Loop Cooling,” IEEE Access, vol. 8, pp. 73 475–

175
References

73 493, 2020.
[10] R. R. Moghaddam, S. Nategh, J. Islam, and A. Boglietti, “Different traction
motor topologies used in e-mobility : Part ii: Magnet-based solutions,” in
Proc. IEEE International Conference on Electrical Machines (ICEM), vol. 1,
2020, pp. 138–143.
[11] B. Bilgin, P. Magne, P. Malysz, Y. Yang, V. Pantelic, M. Preindl, A. Ko-
robkine, W. Jiang, M. Lawford, and A. Emadi, “Making the case for electrified
transportation,” IEEE Trans.Transportation Electrification, vol. 1, no. 1, pp.
4–17, 2015.
[12] J. Islam, S. Nategh, R. R. Moghaddam, and A. Boglietti, “Different Traction
Motor Topologies Used in E-mobility : Part I: Solutions without magnet,”
in Proc. IEEE International Conference on Electrical Machines (ICEM),
vol. 1, 2020, pp. 131–137.
[13] Z. Yang, F. Shang, I. P. Brown, and M. Krishnamurthy, “Comparative
study of interior permanent magnet, induction, and switched reluctance
motor drives for EV and HEV applications,” IEEE Trans. Transportation
Electrification, vol. 1, no. 3, pp. 245–254, 2015.
[14] J. Tang, Y. Liu, and N. Sharma, “Modeling and experimental verification of
high-frequency inductive brushless exciter for electrically excited synchronous
machines,” IEEE Trans. Industry Applications, vol. 55, no. 5, pp. 4613–4623,
2019.
[15] J. Tang, “Synchronous Machines with High-Frequency Brushless Excita-
tion for Vehicle Applications,” Licentiate Thesis, Chalmers University of
Technology, Gothenburg, Sweden, 2019.
[16] A. Wang, C. Wang, and W. L. Soong, “Design and optimization of interior
PM machines with distributed and fractional-slot concentrated-windings
for hybrid electric vehicles,” in Proc. 2014 IEEE Conference and Expo
Transportation Electrification Asia-Pacific (ITEC Asia-Pacific), 2014, pp.
1–5.
[17] X. Fan, B. Zhang, R. Qu, D. Li, J. Li, and Y. Huo, “Comparative thermal
analysis of IPMSMs with integral-slot distributed-winding (ISDW) and
fractional-slot concentrated-winding (FSCW) for electric vehicle application,”
IEEE Trans. Industry Applications, vol. 55, no. 4, pp. 3577–3588, 2019.
[18] G. Berardi, S. Nategh, N. Bianchi, and Y. Thioliere, “A Comparison Between
Random and Hairpin Winding in E-mobility Applications,” in Proc. IECON

176
References

2020 The 46th Annual Conference of the IEEE Industrial Electronics Society.
IEEE, 2020, pp. 815–820.
[19] Q. Ren, D. Crolla, and A. Morris, “Effect of transmission design on electric
vehicle (EV) performance,” in Proc. IEEE vehicle power and propulsion
conference, 2009, pp. 1260–1265.
[20] P. Spanoudakis, N. C. Tsourveloudis, L. Doitsidis, and E. S. Karapidakis,
“Experimental research of transmissions on electric vehicles’ energy consump-
tion,” Energies, vol. 12, no. 3, p. 388, 2019.
[21] W. Wang, Z. Zhang, J. Shi, C. Lin, and Y. Gao, “Optimization of a dual-
motor coupled powertrain energy management strategy for a battery electric
bus based on dynamic programming method,” IEEE Access, vol. 6, pp.
32 899–32 909, 2018.
[22] F. J. Verbruggen, E. Silvas, and T. Hofman, “Electric Powertrain Topology
Analysis and Design for Heavy-Duty Trucks,” Energies, vol. 13, no. 10, p.
2434, 2020.
[23] Tesla, “Model 3,” accessed 2020-11-23. Available at https://www.tesla.com/
en_eu/model3.
[24] K.-J. Tseng and G. Chen, “Computer-aided design and analysis of direct-
driven wheel motor drive,” IEEE Trans.Power Electronics, vol. 12, no. 3, pp.
517–527, 1997.
[25] Y. Zhao, Y. Zhang, and Y. Zhao, “Stability control system for four-in-wheel-
motor drive electric vehicle,” in Proc. IEEE 6th International Conference on
Fuzzy Systems and Knowledge Discovery, vol. 4, 2009, pp. 171–175.
[26] S. E. Schulz, “Exploring the high-power inverter: Reviewing critical design
elements for electric vehicle applications,” IEEE Electrification Magazine,
vol. 5, no. 1, pp. 28–35, March 2017.
[27] J. Reimers, L. Dorn-Gomba, C. Mak, and A. Emadi, “Automotive trac-
tion inverters: Current status and future trends,” IEEE Trans. Vehicular
Technology, vol. 68, no. 4, pp. 3337–3350, 2019.
[28] Volvo Cars, “Volvo XC40,” accessed 2020-11-23. Available at https://www.
volvocars.com/intl/v/cars/xc40-electric.
[29] Polestar, “The Polestar 2,” accessed 2020-11-23. Available at https://www.
polestar.com/cars/polestar-2.
[30] BMW, “The Concept i4,” accessed 2020-11-23. Available at https://www.

177
References

bmw.co.uk/en/topics/discover/concept-cars/bmw-concept-i4.html.
[31] C. Jung, “Power up with 800-v systems: The benefits of upgrading voltage
power for battery-electric passenger vehicles,” IEEE Electrification Magazine,
vol. 5, no. 1, pp. 53–58, March 2017.
[32] IEA Global EV Outlook, “Entering the Decade of Electric Drive,” 2020.
[33] A. Bubert, K. Oberdieck, H. Xu, and R. W. De Doncker, “Experimental
validation of design concepts for future EV-traction inverters,” in Proc.
IEEE Transportation Electrification Conference and Expo (ITEC), 2018, pp.
795–802.
[34] J. Colmenares, “Extreme Implementations of Wide-Bandgap Semiconductors
in Power Electronics,” Ph.D. dissertation, KTH Royal Institute of Technology,
2016.
[35] G. Berardi, S. Nategh, and N. Bianchi, “Inter-turn Voltage in Hairpin
Winding of Traction Motors Fed by High-Switching Frequency Inverters,” in
Proc. IEEE International Conference on Electrical Machines (ICEM), vol. 1,
2020, pp. 909–915.
[36] K. Olejniczak, T. McNutt, D. Simco, A. Wijenayake, T. Flint, B. Passmore,
R. Shaw, D. Martin, A. Curbow, J. Casady, and B. Hull, “A 200 kVA electric
vehicle traction drive inverter having enhanced performance over its entire
operating region,” in Proc. 2017 IEEE 5th Workshop on Wide Bandgap
Power Devices and Applications (WiPDA), Oct 2017, pp. 335–341.
[37] T. J. Hammons, “Tidal power,” Proceedings of the IEEE, vol. 81, no. 3, pp.
419–433, 1993.
[38] T. Hammons, “Tidal power in the United Kingdom,” in Proc. 43rd Interna-
tional Universities Power Engineering Conference, 2008, pp. 1–8.
[39] A. G. Bryans, B. Fox, P. A. Crossley, and M. O’Malley, “Impact of tidal
generation on power system operation in Ireland,” IEEE Trans. Power
Systems, vol. 20, no. 4, pp. 2034–2040, 2005.
[40] A. K. Sikder, N. A. Khan, M. T. Rahman, M. Islam, and A. Hoque, “Tidal
energy: A solution to energy crisis in coastal area of Bangladesh,” in Proc.
International Conference on Electrical Engineering and Information & Com-
munication Technology, 2014, pp. 1–6.
[41] A. S. Bahaj and L. E. Myers, “Fundamentals applicable to the utilisation of
marine current turbines for energy production,” Renewable energy, vol. 28,

178
References

no. 14, pp. 2205–2211, 2003.


[42] G. Mauboussin, “L’usine marémotrice de La Rance,” in Tidal Power.
Springer, 1972, pp. 189–214.
[43] D. J. Olinger and Y. Wang, “Hydrokinetic energy harvesting using tethered
undersea kites,” Journal of Renewable and Sustainable Energy, vol. 7, no. 4,
p. 043114, 2015.
[44] Minesto AB, “The future of renewable energy [Online],” accessed 2020-11-23.
Available at https://minesto.com/our-technology.
[45] M. Landberg, “Submersible plant,” Aug. 21 2012, US Patent 8,246,293.
[46] “Power Take-Off System for a Subsea Tidal Kite [Online],” accessed 2020-11-
23. Available at https://www.powerkite-project.eu/.
[47] M. L. Loyd, “Crosswind kite power (for large-scale wind power production),”
Journal of energy, vol. 4, no. 3, pp. 106–111, 1980.
[48] P. Bechtle, T. Gehrmann, C. Sieg, and U. Zillmann, “Awesome: An open-
source test platform for airborne wind energy systems,” arXiv preprint
arXiv:1704.08695, 2017.
[49] C. Jehle, “Automatic flight control of tethered kites for power generation,”
Diploma Thesis, Technical University Munich, Germany, 2012.
[50] S. Gros and M. Diehl, “Modeling of airborne wind energy systems in natural
coordinates,” in Airborne wind energy. Springer, 2013, pp. 181–203.
[51] J. W. Kolar, T. Friedli, F. Krismer, A. Looser, M. Schweizer, R. A. Friede-
mann, P. K. Steimer, and J. B. Bevirt, “Conceptualization and multiobjective
optimization of the electric system of an airborne wind turbine,” IEEE Jour-
nal of Emerging and Selected Topics in Power Electronics, vol. 1, no. 2, pp.
73–103, 2013.
[52] M. S. Ahmed, A. Hably, and S. Bacha, “Kite generator system modeling
and grid integration,” IEEE Trans. Sustainable Energy, vol. 4, no. 4, pp.
968–976, 2013.
[53] E. C. Malz, “Airborne wind energy–to fly or not to fly?” Doctoral Thesis,
Chalmers University of Technology, Gothenburg, Sweden, 2020.
[54] E. Singhroy, “Improving power performance of tethered turbine tidal farms,”
Master’s Thesis, Energy and Environment Dept., Chalmers Univ., Gothen-
burg, Sweden, 2017.

179
References

[55] H. M. Kojabadi, L. Chang, and T. Boutot, “Development of a novel wind


turbine simulator for wind energy conversion systems using an inverter-
controlled induction motor,” IEEE Trans. Energy Conversion, vol. 19, no. 3,
pp. 547–552, 2004.
[56] F. Martinez, L. C. Herrero, and S. de Pablo, “Open loop wind turbine
emulator,” Renewable Energy, vol. 63, pp. 212–221, 2014.
[57] D. S. Dolan and P. Lehn, “Real-Time Wind Turbine Emulator Suitable
for Power Quality and Dynamic Control Studies,” in Proc. International
Conference on Power Systems Transients (IPST’05), 2005.
[58] P.-Y. Chen, K.-W. Hu, Y.-G. Lin, and C.-M. Liaw, “Development of a prime
mover emulator using a permanent-magnet synchronous motor drive,” IEEE
Trans. Power Electronics, vol. 33, no. 7, pp. 6114–6125, 2018.
[59] L. Lopes, J. Lhuilier, M. Khokar, and A. Mukherjee, “A wind turbine
emulator that represents the dynamics of the wind turbine rotor and drive
train,” in Proc. IEEE 36th Power Electronics Specialists Conference, 2005,
pp. 2092–2097.
[60] C. Wei, Z. Zhang, W. Qiao, and L. Qu, “An adaptive network-based reinforce-
ment learning method for MPPT control of PMSG wind energy conversion
systems,” IEEE Trans. Power Electronics, vol. 31, no. 11, pp. 7837–7848,
2016.
[61] M. N. Uddin and N. Patel, “Maximum power point tracking control of
IPMSG incorporating loss minimization and speed sensorless schemes for
wind energy system,” IEEE Trans. Industry Applications, vol. 52, no. 2, pp.
1902–1912, 2016.
[62] T. Nakamura, S. Morimoto, M. Sanada, and Y. Takeda, “Optimum control of
IPMSG for wind generation system,” in Proc. Power Conversion Conference-
Osaka (Cat. No. 02TH8579), vol. 3, 2002, pp. 1435–1440.
[63] S. Morimoto, H. Nakayama, M. Sanada, and Y. Takeda, “Sensorless out-
put maximization control for variable-speed wind generation system using
IPMSG,” IEEE Trans. Industry Applications, vol. 41, no. 1, pp. 60–67, 2005.
[64] Z. Zhang, Y. Zhao, W. Qiao, and L. Qu, “A space-vector-modulated sensorless
direct-torque control for direct-drive PMSG wind turbines,” IEEE Trans.
Industry Applications, vol. 50, no. 4, pp. 2331–2341, 2014.
[65] “IEEE Guide for the Functional Specification of Medium Voltage (1 kV - 35

180
References

kV) Electronic Shunt Devices for Dynamic Voltage Compensation,” IEEE


Std 1623-2004, pp. 1–23, 2005.
[66] A. Massoud, S. Finney, and B. Williams, “Multilevel converters and series
connection of IGBT evaluation for high-power, high-voltage applications,”
in Proc. Second International Conference on Power Electronics, Machines
and Drives (PEMD 2004), vol. 1, 2004, pp. 1–5.
[67] G. Wang, G. Konstantinou, C. D. Townsend, J. Pou, S. Vazquez, G. D.
Demetriades, and V. G. Agelidis, “A review of power electronics for grid
connection of utility-scale battery energy storage systems,” IEEE Trans.
Sustainable Energy, vol. 7, no. 4, pp. 1778–1790, 2016.
[68] G.-J. Su, “Multilevel DC-link inverter,” IEEE Trans. Industry Applications,
vol. 41, no. 3, pp. 848–854, 2005.
[69] E. Najafi and A. H. M. Yatim, “Design and implementation of a new multilevel
inverter topology,” IEEE Trans. Industrial Electronics, vol. 59, no. 11, pp.
4148–4154, 2012.
[70] O. Josefsson, T. Thiringer, S. Lundmark, and H. Zelaya, “Evaluation and
comparison of a two-level and a multilevel inverter for an EV using a
modulized battery topology,” in Proc. IECON-38th Annual Conference on
IEEE Industrial Electronics Society, 2012, pp. 2949–2956.
[71] V. Naumanen, J. Korhonen, P. Silventoinen, and J. Pyrhönen, “Mitigation
of high du/dt-originated motor overvoltages in multilevel inverter drives,”
IET Power Electronics, vol. 3, no. 5, pp. 681–689, 2010.
[72] H. Zhang, S. Finney, T. Lim, A. Massoud, J. Yang, and B. Williams, “DC-link
capacitor voltage balancing for a five-level diode-clamped active power filter
using redundant vectors,” Australian Journal of Electrical and Electronics
Engineering, vol. 10, no. 2, pp. 137–147, 2013.
[73] S. Guanchu, K. Lee, L. Xinchun, and L. Chongbo, “New neutral point
balancing strategy for five-level diode clamped converters used in STATCOM
of wind energy conversion systems,” in Proc. IEEE 6th International Power
Electronics and Motion Control Conference, 2009, pp. 2354–2358.
[74] M. Mazuela, I. Baraia, A. Sanchez-Ruiz, I. Echeverria, I. Torre, and I. Atutxa,
“DC-link voltage balancing strategy based on SVM and reactive power
exchange for a 5l-MPC back-to-back converter for medium-voltage drives,”
IEEE Trans. Industrial Electronics, vol. 63, no. 12, pp. 7864–7875, 2016.

181
References

[75] M. Saeedifard, R. Iravani, and J. Pou, “A space vector modulation strategy for
a back-to-back five-level HVDC converter system,” IEEE Trans. Industrial
Electronics, vol. 56, no. 2, pp. 452–466, 2009.
[76] P. Qashqai, A. Sheikholeslami, H. Vahedi, and K. Al-Haddad, “A new SVM-
based voltage balancing method for five-level NPC inverter,” in Proc. 7th
Power Electronics and Drive Systems Technologies Conference (PEDSTC),
2016, pp. 511–516.
[77] F. Z. Peng, Z. Pan, K. Corzine, V. Stefanovic, and M. Leuthen, “Voltage
balancing control of diode-clamped multilevel rectifier/inverter systems,”
IEEE Trans. Industry Applications, vol. 41, no. 6, pp. 1698–1706, 2005.
[78] W. Yue, L. Ning, L. Su, C. Wulong, L. Wanjun, and W. Zhao’an, “Research on
DC capacitor voltage self-balancing space vector modulation strategy of five-
level npc converter,” in Proc. IEEE Applied Power Electronics Conference
and Exposition (APEC’14), 2014, pp. 2694–2699.
[79] H. Hotait, A. Massoud, S. Finney, and B. Williams, “Capacitor voltage
balancing using redundant states of space vector modulation for five-level
diode clamped inverters,” IET Power Electronics, vol. 3, no. 2, pp. 292–313,
2010.
[80] H. Zhang, S. J. Finney, A. Massoud, and B. W. Williams, “An SVM algorithm
to balance the capacitor voltages of the three-level NPC active power filter,”
IEEE Trans. Power Electronics, vol. 23, no. 6, pp. 2694–2702, 2008.
[81] G. P. Adam, S. J. Finney, A. M. Massoud, and B. W. Williams, “Capacitor
balance issues of the diode-clamped multilevel inverter operated in a quasi
two-state mode,” IEEE Transactions on Industrial Electronics, vol. 55, no. 8,
pp. 3088–3099, 2008.
[82] G. P. Adam, S. Finney, O. Ojo, and B. Williams, “Quasi-two-level and three-
level operation of a diode-clamped multilevel inverter using space vector
modulation,” IET Power Electronics, vol. 5, no. 5, pp. 542–551, 2012.
[83] C. Newton and M. Sumner, “Novel technique for maintaining balanced inter-
nal DC link voltages in diode clamped five-level inverters,” IEE Proceedings-
Electric Power Applications, vol. 146, no. 3, pp. 341–349, 1999.
[84] N. Hatti, Y. Kondo, and H. Akagi, “Five-level diode-clamped PWM converters
connected back-to-back for motor drives,” IEEE Trans. Industry Applications,
vol. 44, no. 4, pp. 1268–1276, 2008.

182
References

[85] A. Rufer, N. Koch, and N. Cherix, “Control of the actively balanced capacitive
voltage divider for a five-level NPC inverter-estimation of the intermediary
levels currents,” in Proc. PCIM Europe; International Exhibition and Con-
ference for Power Electronics, Intelligent Motion, Renewable Energy and
Energy Management, 2016, pp. 1–7.
[86] Z. Shu, H. Zhu, X. He, N. Ding, and Y. Jing, “One-inductor-based auxiliary
circuit for dc-link capacitor voltage equalisation of diode-clamped multilevel
converter,” IET Power Electronics, vol. 6, no. 7, pp. 1339–1349, 2013.
[87] K. Sano and H. Fujita, “Voltage-balancing circuit based on a resonant
switched-capacitor converter for multilevel inverters,” IEEE Trans. Industry
Applications, vol. 44, no. 6, pp. 1768–1776, 2008.
[88] D. Cui and Q. Ge, “A novel hybrid voltage balance method for five-level
diode-clamped converters,” IEEE Trans. Industrial Electronics, vol. 65, no. 8,
pp. 6020–6031, 2018.
[89] T. Lagier, P. Ladoux, and P. Dworakowski, “Potential of silicon carbide
MOSFETs in the DC/DC converters for future HVDC offshore wind farms,”
IET High voltage, vol. 2, no. 4, pp. 233–243, 2017.
[90] L. A. Vitoi, D. I. Brandao, and E. Tedeschi, “Power quality enhancement by
SiC Active Power Filters in Oil and Gas Platforms,” in Proc. IEEE Energy
Conversion Congress and Exposition (ECCE), 2019, pp. 4299–4304.
[91] F. Blaabjerg and K. Ma, “Future on power electronics for wind turbine sys-
tems,” IEEE Journal of Emerging and Selected Topics in Power Electronics,
vol. 1, no. 3, pp. 139–152, 2013.
[92] L. Harnefors, Control of variable-speed drives. Applied Signal Processing
and Control, Department of Electronics, Mälardalen University, 2002.
[93] L. Harnefors and H.-P. Nee, “A general algorithm for speed and position
estimation of AC motors,” IEEE Trans. Industrial Electronics, vol. 47, no. 1,
pp. 77–83, 2000.
[94] S. Morimoto, Y. Takeda, T. Hirasa, and K. Taniguchi, “Expansion of operat-
ing limits for permanent magnet motor by current vector control considering
inverter capacity,” IEEE Trans. Industry Applications, vol. 26, no. 5, pp.
866–871, 1990.
[95] F.-J. Lin, Y.-C. Hung, J.-M. Chen, and C.-M. Yeh, “Sensorless IPMSM
drive system using saliency back-EMF-based intelligent torque observer with

183
References

MTPA control,” IEEE Trans. Industrial Informatics, vol. 10, no. 2, pp.
1226–1241, 2014.
[96] K. Li and Y. Wang, “Maximum torque per ampere (MTPA) control for
IPMSM drives using signal injection and an MTPA control law,” IEEE
Trans. Industrial Informatics, vol. 15, no. 10, pp. 5588–5598, 2019.
[97] C. Mademlis, I. Kioskeridis, and N. Margaris, “Optimal efficiency control
strategy for interior permanent-magnet synchronous motor drives,” IEEE
Trans. Energy Conversion, vol. 19, no. 4, pp. 715–723, 2004.
[98] A. Balamurali, G. Feng, A. Kundu, H. Dhulipati, and N. C. Kar, “Noninvasive
and Improved Torque and Efficiency Calculation Toward Current Advance
Angle Determination for Maximum Efficiency Control of PMSM,” IEEE
Trans. Transportation Electrification, vol. 6, no. 1, pp. 28–40, 2019.
[99] S. Kim, Y.-D. Yoon, S.-K. Sul, and K. Ide, “Maximum torque per ampere
(MTPA) control of an IPM machine based on signal injection considering
inductance saturation,” IEEE Trans. Power Electronics, vol. 28, no. 1, pp.
488–497, 2012.
[100] L. Tian, Y. Liu, J. Zhao, and J. Sun, “A novel MTPA control for IPMSMs
based on extremum seeking control,” in Proc. 34th Chinese Control Confer-
ence (CCC), 2015, pp. 4268–4273.
[101] L. Angquist and M. Bongiorno, “Auto-normalizing phase-locked loop for
grid-connected converters,” in Proc. IEEE Energy Conversion Congress and
Exposition, 2009, pp. 2957–2964.
[102] A. Yazdani and R. Iravani, Voltage-sourced converters in power systems:
modeling, control, and applications. John Wiley & Sons, 2010.
[103] Y. Wang, J. Meng, X. Zhang, and L. Xu, “Control of PMSG-based wind
turbines for system inertial response and power oscillation damping,” IEEE
Trans. Sustainable Energy, vol. 6, no. 2, pp. 556–574, 2015.
[104] X. Zhao, Z. Yan, Y. Xue, and X.-P. Zhang, “Wind power smoothing by
controlling the inertial energy of turbines with optimized energy yield,” IEEE
Access, vol. 5, pp. 23 374–23 382, 2017.
[105] X. Zhao, Z. Yan, and X.-P. Zhang, “A wind-wave farm system with self-energy
storage and smoothed power output,” IEEE Access, vol. 4, pp. 8634–8642,
2016.
[106] M. AB, “Holyhead Deep - the world’s first low-flow tidal stream project

184
References

[Online],” accessed 2020-11-23. Available at https://minesto.com/projects/


holyhead-deep.
[107] W. Batten, A. Bahaj, A. Molland, and J. Chaplin, “The prediction of the
hydrodynamic performance of marine current turbines,” Renewable energy,
vol. 33, no. 5, pp. 1085–1096, 2008.
[108] Z. Zhou, F. Scuiller, J. F. Charpentier, M. E. H. Benbouzid, and T. Tang,
“Power control of a nonpitchable PMSG-based marine current turbine at
overrated current speed with flux-weakening strategy,” IEEE Journal of
Oceanic Engineering, vol. 40, no. 3, pp. 536–545, 2015.
[109] G. Caraiman, C. Nichita, V. Mînzu, and B. Dakyo, “Marine current turbine
simulator development based on hardware in the loop simulation concept,”
Journal of Energy and Power Engineering, vol. 5, no. 9, 2011.
[110] K. Clark, N. W. Miller, and J. J. Sanchez-Gasca, “Modeling of GE wind
turbine-generators for grid studies,” GE energy, vol. 4, pp. 4–13, 2010.
[111] R. Cárdenas, R. Peña, S. Alepuz, and G. Asher, “Overview of control
systems for the operation of DFIGs in wind energy applications,” IEEE
Trans. Industrial Electronics, vol. 60, no. 7, pp. 2776–2798, 2013.
[112] C. Mademlis, Σερβoκινητ ήρια Συστ ήµατ α[Servomotor Systems]. Thes-
saloniki, Greece: Tziolas Publications, 2010.
[113] ABB, “ACS 600 MultiDrive Manuals [Online],” accessed 2020-11-23. Available
at https://library.e.abb.com/public/fdc7fc079eb9493ac2256def0041557b/en_
ACA6xxhw_1e_scrres.pdf.
[114] R. Marquardt, “Stromrichterschaltungen mit verteilten energiespeichern,”
2001, German Patent DE10103031A1.
[115] Y. Liang and C. Nwankpa, “A new type of STATCOM based on cascading
voltage-source inverters with phase-shifted unipolar SPWM,” IEEE Trans.
Industry Applications, vol. 35, no. 5, pp. 1118–1123, 1999.
[116] B. Gemmell, J. Dorn, D. Retzmann, and D. Soerangr, “Prospects of multilevel
VSC technologies for power transmission,” in Proc. IEEE/PES Transmission
and Distribution Conference and Exposition, 2008, pp. 1–16.
[117] V. Yaramasu, B. Wu, P. C. Sen, S. Kouro, and M. Narimani, “High-power
wind energy conversion systems: State-of-the-art and emerging technologies,”
Proceedings of the IEEE, vol. 103, no. 5, pp. 740–788, 2015.
[118] R. H. Baker, “High-voltage converter circuit,” May 13 1980, US Patent

185
References

4,203,151.
[119] ABB, “Medium voltage AC drive ACS 6000 [Online],” ac-
cessed 2020-11-23. Available at https://library.e.abb.com/public/
5e7f2eb1c4d429c948257e12004d4282/ACS%206000%20EN%20Rev%
20F_lowres.pdf.
[120] Danfoss, “Vacon-3000 [Online],” accessed 2020-11-23. Available at
http://danfoss.ipapercms.dk/Drives/DD/Global/SalesPromotion/
Factsheets/ProductFactsheets/UK/vacon-3000/.
[121] S. Kouro, M. Malinowski, K. Gopakumar, J. Pou, L. G. Franquelo, B. Wu,
J. Rodriguez, M. A. Pérez, and J. I. Leon, “Recent advances and industrial
applications of multilevel converters,” IEEE Trans. Industrial Electronics,
vol. 57, no. 8, pp. 2553–2580, 2010.
[122] A. Sanchez-Ruiz, M. Mazuela, S. Alvarez, G. Abad, and I. Baraia, “Medium
voltage–high power converter topologies comparison procedure, for a 6.6
kv drive application using 4.5 kv IGBT modules,” IEEE Trans. Industrial
Electronics, vol. 59, no. 3, pp. 1462–1476, 2012.
[123] J. Rodríguez, S. Bernet, B. Wu, J. O. Pontt, and S. Kouro, “Multilevel
voltage-source-converter topologies for industrial medium-voltage drives,”
IEEE Trans. Industrial Electronics, vol. 54, no. 6, pp. 2930–2945, 2007.
[124] T. Bruckner, S. Bernet, and H. Guldner, “The active NPC converter and its
loss-balancing control,” IEEE Trans. Industrial Electronics, vol. 52, no. 3,
pp. 855–868, 2005.
[125] G. Carrara, S. Gardella, M. Marchesoni, R. Salutari, and G. Sciutto, “A
new multilevel PWM method: A theoretical analysis,” IEEE Trans. Power
Electronics, vol. 7, no. 3, pp. 497–505, 1992.
[126] B. P. McGrath and D. G. Holmes, “Multicarrier PWM strategies for multilevel
inverters,” IEEE Trans. Industrial Electronics, vol. 49, no. 4, pp. 858–867,
2002.
[127] S. Wei, B. Wu, F. Li, and C. Liu, “A general space vector PWM control
algorithm for multilevel inverters,” in Proc. IEEE Applied Power Electronics
Conference and Exposition (APEC’03), vol. 1, 2003, pp. 562–568.
[128] S. K. Mondal, B. K. Bose, V. Oleschuk, and J. O. Pinto, “Space vector
pulse width modulation of three-level inverter extending operation into
overmodulation region,” IEEE Trans. Power Electronics, vol. 18, no. 2, 2003.

186
References

[129] J. Pou, R. Pindado, and D. Boroyevich, “Voltage-balance limits in four-


level diode-clamped converters with passive front ends,” in Proc. IEEE 28th
Annual Conference of the Industrial Electronics Society (IECON’02), vol. 2,
2002, pp. 898–902.
[130] X.-j. Zeng, H. Wang, X. Zhang, D.-l. Gong, and X. Yang, “Transformer-less
five-level high-voltage full-scale converter with a self-balance auxiliary circuit
for 10 MW wind turbines,” IET Power Electronics, vol. 10, no. 14, pp.
1841–1851, 2017.
[131] A. Rufer, “A five-level NPC photovoltaic inverter with an actively balanced
capacitive voltage divider,” in Proc. PCIM Europe; International Exhibition
and Conference for Power Electronics, Intelligent Motion, Renewable Energy
and Energy Management, 2015, pp. 1–8.
[132] G. Mademlis, G. K. Steinke, and A. Rufer, “Feed-forward-based control in a
dc–dc converter of asymmetric multistage-stacked boost architecture,” IEEE
Trans. Power Electronics, vol. 32, no. 2, pp. 1507–1517, 2017.
[133] J.-J. Huselstein, C. Gauthier, and C. Glaize, “Use of the mosfet channel
reverse conduction in an inverter for suppression of the integral diode recov-
ery current,” in Proc. 5th European Conference on Power Electronics and
Applications, 1993, pp. 431–436.
[134] Y. Jiao, F. C. Lee, and S. Lu, “Space vector modulation for three-level NPC
converter with neutral point voltage balance and switching loss reduction,”
IEEE Trans. Power Electronics, vol. 29, no. 10, pp. 5579–5591, 2014.
[135] A. Elasser and T. P. Chow, “Silicon carbide benefits and advantages for
power electronics circuits and systems,” Proceedings of the IEEE, vol. 90,
no. 6, pp. 969–986, 2002.
[136] J. Biela, M. Schweizer, S. Waffler, and J. W. Kolar, “SiC versus Si - evaluation
of potentials for performance improvement of inverter and DC–DC converter
systems by SiC power semiconductors,” IEEE Trans. Industrial Electronics,
vol. 58, no. 7, pp. 2872–2882, 2011.
[137] Wolfspeed Power & RF, “Cree SiC Technology Enables the Quan-
tum Reduction of Solar Power Inverter Size, Weight & Cost [On-
line],” accessed 2020-11-23. Available at https://www.wolfspeed.com/news/
sic-mosfets-for-solar-inverters.
[138] Texas Instruments, “Fundamentals of MOSFET and IGBT gate driver circuits
[Online],” 2017, accessed 2020-11-23. Available at http://www.ti.com/lit/

187
References

ml/slua618a/slua618a.pdf.
[139] ——, “PCB Design Guidelines For Reduced EMI [Online],” 1999, accessed
2020-11-23. Available at http://www.ti.com/lit/an/szza009/szza009.pdf.
[140] SCH2080KE, ROHM Semiconductor, October 2018, rev. 003.
[141] R. Tarzwell and K. Bahl, “High voltage printed circuit design & manufactur-
ing notebook,” SIERRA Proto Express, Tech. Rep., 2004.
[142] R. Tarzwell, “High power printed circuit design notebook,” SIERRA Proto
Express, Tech. Rep., 2004.
[143] Isola, “370HR High Performance Laminate and Prepreg [Online],” 2016,
accessed 2020-11-23. Available at https://www.isola-group.com/wp-content/
uploads/data-sheets/370hr.pdf.
[144] Texas Instruments, “TMS320F2837xD Dual-Core Delfino Microcontrollers
Technical Reference Manual [Online],” 2013, accessed 2020-11-23. Available
at http://www.ti.com/lit/ug/spruhm8h/spruhm8h.pdf.
[145] S. Yamamoto, T. Kano, Y. Yamaguchi, and T. Ara, “A method to determine
direct-and quadrature-axis inductances of permanent magnet synchronous
motors,” Electrical Engineering in Japan, vol. 171, no. 3, pp. 41–50, 2010.
[146] M. Kazerooni and N. C. Kar, “Methods for determining the parameters and
characteristics of PMSM,” in Proc. IEEE International Electric Machines &
Drives Conference (IEMDC), 2011, pp. 955–960.
[147] K. S. Amitkumar, R. S. Kaarthik, and P. Pillay, “A Versatile Power-Hardware-
in-the-Loop-Based Emulator for Rapid Testing of Transportation Electric
Drives,” IEEE Trans. Transportation Electrification, vol. 4, no. 4, pp. 901–
911, 2018.
[148] O. Mohammed, Z. Liu, and S. Liu, “Equivalent hardware representation of
PM synchronous machines realized by a physics-based machine model for
hardware-in-the-loop simulation applications,” in Proc. 2008 34th Annual
Conference of IEEE Industrial Electronics, 2008, pp. 1185–1188.
[149] Y. Luo, M. Awal, W. Yu, and I. Husain, “FPGA Based High Bandwidth
Motor Emulator for Interior Permanent Magnet Machine Utilizing SiC
Power Converter,” IEEE Journal of Emerging and Selected Topics in Power
Electronics, 2020.
[150] K. Saito and H. Akagi, “A Power Hardware-in-the-Loop (P-HIL) Test Bench
Using Two Modular Multilevel DSCC Converters for a Synchronous Motor

188
References

Drive,” IEEE Trans. on Industry Applications, vol. 54, no. 5, pp. 4563–4573,
2018.
[151] ——, “A Real-Time Real-Power DSCC-Based Emulator Capable of Repro-
ducing Both Bending and Torsional Vibrations of a Motor and Load,” in
Proc. 2019 IEEE Energy Conversion Congress and Exposition (ECCE), 2019,
pp. 306–312.
[152] J. Itoh and T. Abe, “Circulation current reduction for a motor simulator
system using a power converter with a common mode transformer,” in
Proc. 2009 IEEE 6th International Power Electronics and Motion Control
Conference, 2009, pp. 1662–1667.
[153] L. Rovere, A. Formentini, G. L. Calzo, P. Zanchetta, and T. Cox, “Zero-
Sequence Voltage Elimination for Dual-Fed Common DC-Link Open-End
Winding PMSM High-Speed Starter-Generator-Part I: Modulation,” IEEE
Trans. Industry Applications, vol. 55, no. 6, pp. 7804–7812, 2019.
[154] K. Saito and H. Akagi, “A Real-Time Real-Power Emulator of a Medium-
Voltage High-Speed Induction Motor Coupled with a Mechanical Load,” in
Proc. 2018 IEEE Energy Conversion Congress and Exposition (ECCE), 2018,
pp. 5242–5249.
[155] Z. Liu, O. A. Mohammed, and S. Liu, “Equivalent hardware representation
of PM synchronous motors from the physics-based phase variable model
obtained through FE computation,” IEEE Trans. Magnetics, vol. 45, no. 3,
pp. 1450–1453, 2009.
[156] H. Abu-Rub, M. Malinowski, and K. Al-Haddad, “Hardware-in-the-Loop
Systems with Power Electronics: A Powerful Simulation Tool,” in Power
Electronics for Renewable Energy Systems, Transportation and Industrial
Applications. Wiley-IEEE Press, 2014, pp. 573–590.
[157] Y. Gao and J. Ye, “A novel control method of zero sequence circulation in
parallel inverters,” in Proc. 2019 International Conference on Intelligent
Transportation, Big Data Smart City (ICITBS), 2019, pp. 465–468.
[158] W. Jiang, W. Ma, J. Wang, W. Wang, X. Zhang, and L. Wang, “Suppres-
sion of Zero Sequence Circulating Current for Parallel Three-Phase Grid-
Connected Converters Using Hybrid Modulation Strategy,” IEEE Trans.
Industrial Electronics, vol. 65, no. 4, pp. 3017–3026, 2018.
[159] Wolfspeed Power & RF, “XM3 Half-Bridge Power Module Family [On-
line],” accessed 2020-11-23. Available at https://www.wolfspeed.com/

189
References

xm3-power-module-family.
[160] EPA, “Vehicle Weight Classifications for the Emission Stan-
dards Reference Guide,” accessed 2020-11-23. Available
at https://www.epa.gov/emission-standards-reference-guide/
vehicle-weight-classifications-emission-standards-reference-guide.
[161] European Commission, “Vehicle categories,” accessed 2020-11-23. Available
at https://ec.europa.eu/growth/sectors/automotive/vehicle-categories_en.
[162] E. Grunditz, “Design and assessment of battery electric vehicle powertrain,
with respect to performance, energy consumption and electric motor thermal
capability,” Doctoral Thesis, Chalmers University of Technology, Gothenburg,
Sweden, 2016.
[163] E. Arfa Grunditz and T. Thiringer, “Electric vehicle igbt power module
sizing and drive cycle energy efficiency for various switching frequencies-
based on a scalable module model,” in Proc. 2018 20th European Conference
on Power Electronics and Applications (EPE’18 ECCE Europe), Sep. 2018,
pp. P.1–P.10.
[164] SAE, “Electric Vehicle Acceleration, Gradeability, and Deceleration Test
Procedure,” SAE Ground Vehicle Standard J1666_199305, 1993.
[165] “Test Specification Acceleration and Deceleration Testing,” Center for
Evaluation of Clean Energy Technology (CECET), Tech. Rep., 2015,
accessed 2020-11-23. [Online]. Available: https://avt.inl.gov/sites/default/
files/pdf/reports/IntertekTestSpecAccelDecelTesting.pdf
[166] M. Tutuianu, P. Bonnel, B. Ciuffo, T. Haniu, N. Ichikawa, A. Marotta,
J. Pavlovic, and H. Steven, “Development of the World-wide harmonized
Light duty Test Cycle (WLTC) and a possible pathway for its introduction
in the European legislation,” Transportation research part D: transport and
environment, vol. 40, pp. 61–75, 2015.
[167] M. André, “The artemis european driving cycles for measuring car pollutant
emissions,” Science of the Total Environment, vol. 334, pp. 73–84, 2004.
[168] Standardization Administration of China, “China automotive test cycle—Part
1: Light-duty vehicles GB/T 38146.1-2019.”
[169] Z. Wu, D. Depernet, C. Kieffer, F. Dubas, D. Hissel, and C. Espanet,
“Electrical motor design for hybrid heavy-duty electrical powertrain,” in
Proc. 2009 IEEE Vehicle Power and Propulsion Conference, Sep. 2009, pp.

190
References

486–493.
[170] H. Stenvall, “Driving resistance analysis of long haulage trucks at Volvo,”
Master’s Thesis, Chalmers University of Technology, 2010.
[171] R. Ramos, “Film capacitors in power applications: Choices and particular
characteristics needed,” IEEE Power Electronics Magazine, vol. 5, no. 1, pp.
45–50, 2018.
[172] H. Wen, W. Xiao, and X. Wen, “Comparative evaluation of DC-link capacitors
for electric vehicle application,” in Proc. IEEE International Symposium on
Industrial Electronics, 2012, pp. 1472–1477.
[173] J.-S. Lai, H. Kouns, and J. Bond, “A low-inductance DC bus capacitor
for high power traction motor drive inverters,” in Proc. IEEE Industry
Applications Conference. 37th IAS Annual Meeting (Cat. No. 02CH37344),
vol. 2, 2002, pp. 955–962.
[174] J. Stewart, J. Neely, J. Delhotal, and J. Flicker, “DC link bus design for
high frequency, high temperature converters,” in Proc. IEEE Applied Power
Electronics Conference and Exposition (APEC), 2017, pp. 809–815.
[175] R. Alizadeh, M. Schupbach, T. Adamson, J. C. Balda, Y. Zhao, S. Long,
K. W. Jung, C. R. Kharangate, M. Asheghi, and K. E. Goodson, “Busbar
Design for Distributed DC-Link Capacitor Banks for Traction Applications,”
in Proc. IEEE Energy Conversion Congress and Exposition (ECCE), 2018,
pp. 4810–4815.
[176] J. W. Kolar and S. D. Round, “Analytical calculation of the RMS current
stress on the DC-link capacitor of voltage-PWM converter systems,” IEE
Proceedings - Electric Power Applications, vol. 153, no. 4, pp. 535–543, 2006.
[177] X. Wen, T. Fan, P. Ning, and Q. Guo, “Technical approaches towards ultra-
high power density SiC inverter in electric vehicle applications,” CES Trans.
Electrical Machines and Systems, vol. 1, no. 3, pp. 231–237, 2017.
[178] KEMET, “Power Film Capacitors - C4E Rectangular Brick
Type DC Link Series,” 2016, accessed 2020-11-23. Available at
https://sh.kemet.com/Lists/ApplicationBriefs/Attachments/138/C4E%
20Rectangular%20Brick%20Type%20DC%20Link%20Series.pdf.
[179] muRata, “What are impedance/ ESR frequency characteristics in capacitors?
[Online],” 2013, accessed 2021-01-21. Available at https://article.murata.
com/en-eu/article/impedance-esr-frequency-characteristics-in-capacitors.

191
References

[180] ABB, “Technical Guide No. 102: Effects of AC Drives on Motor Insulation
[Online],” accessed 2020-11-23. Available at https://library.e.abb.com/public/
fec1a7b62d273351c12571b60056a0fd/voltstress.pdf.
[181] H. S. Seo and D. Shin, “Development of Parallel and Direct Cooling System
for EV/FCEV Inverter,” SAE Technical Paper, 2018.
[182] A. Carriero, M. Locatelli, K. Ramakrishnan, G. Mastinu, and M. Gobbi, “A
review of the state of the art of electric traction motors cooling techniques,”
SAE Technical Paper, 2018.
[183] C. Van Godbold, V. A. Sankaran, and J. L. Hudgins, “Thermal analysis of
high-power modules,” IEEE Trans. Power Electronics, vol. 12, no. 1, pp.
3–11, 1997.
[184] C. Chen, F. Luo, and Y. Kang, “A review of sic power module packaging:
Layout, material system and integration,” CPSS Transactions on Power
Electronics and Applications, vol. 2, no. 3, pp. 170–186, 2017.
[185] Infineon, “Easy IGBT Power Modules [Online],” accessed 2021-01-
20. Available at https://www.infineon.com/cms/en/product/power/igbt/
igbt-modules/easy/.
[186] Wolfspeed Power & RF, “Wolfspeed WolfPACK™ SiC Power Modules Family
[Online],” accessed 2021-01-20. Available at https://www.wolfspeed.com/
wolfspeed-wolfpack-sic-power-modules-family.
[187] Baginski, Patrick, “The module without a baseplate: A reliable and cost-
effective solution [Online],” accessed 2021-01-20. Available at https://m.eet.
com/media/1158352/c0995vincotech.pdf.
[188] K. Namiki, K. Murota, and M. Shoji, “High performance motor and inverter
system for a newly developed electric vehicle,” SAE Technical Paper, 2018.
[189] H. Baumann, P. Heinemeyer, W. Staiger, M. Topfer, K. Unger, and D. Muller,
“Optimized cooling systems for high-power semiconductor devices,” IEEE
Trans. Industrial Electronics, vol. 48, no. 2, pp. 298–306, April 2001.
[190] K. Nakatsu and R. Saito, “The next-generation high power density inverter
technology for vehicle,” in Proc. 2014 International Power Electronics Con-
ference (IPEC-Hiroshima 2014 - ECCE ASIA), May 2014, pp. 1925–1928.
[191] K. Olesen, R. Bredtmann, and R. Eisele, “"ShowerPower" New Cooling
Concept for Automotive Applications,” in Proc. of the Automotive Power
Electronics Conference, Paris, France, June, 2006, pp. 21–22.

192
References

[192] M. Reeves, J. Moreno, P. Beucher, S.-J. Loong, and D. Brown, “Investigation


on the impact on thermal performances of new pin and fin geometries applied
to liquid cooling of power electronics,” in Proc. PCIM Europe, 2011, pp.
772–778.
[193] J. Zhang, X. Zhu, M. E. Mondejar, and F. Haglind, “A review of heat
transfer enhancement techniques in plate heat exchangers,” Renewable and
Sustainable Energy Reviews, vol. 101, pp. 305–328, 2019.
[194] N. Karras, T. Kuthada, and J. Wiedemann, “An approach for water jacket
flow simulations,” SAE Technical Paper, 2014.
[195] Z. Zhou, M. S. Kanniche, S. G. Butcup, and P. Igic, “High-speed electro-
thermal simulation model of inverter power modules for hybrid vehicles,”
IET Electric Power Applications, vol. 5, no. 8, pp. 636–643, Sep. 2011.
[196] A. Kersten, E. Grunditz, and T. Thiringer, “Efficiency of Active Three-
Level and Five-Level NPC Inverters Compared to a Two-Level Inverter in a
Vehicle,” in Proc. 2018 20th European Conference on Power Electronics and
Applications (EPE’18 ECCE Europe), Sep. 2018, pp. P.1–P.9.
[197] C. Tang and T. Thiringer, “Thermal modelling of a mutlichip IGBT power
module,” in Proc. 2019 21st European Conference on Power Electronics and
Applications (EPE ’19 ECCE Europe), Sep. 2019, pp. P.1–P.8.
[198] Y. Song and B. Wang, “Evaluation Methodology and Control Strategies
for Improving Reliability of HEV Power Electronic System,” IEEE Trans.
Vehicular Technology, vol. 63, no. 8, pp. 3661–3676, Oct 2014.
[199] K. Haehndel, A. Pere, T. Frank, F. Christel, and S. Abanteriba, “A numerical
investigation of dampening dynamic profiles for the application in transient
vehicle thermal management simulations,” SAE Technical Paper, 2014.
[200] H. Reister and W. Bauer, “Simulation process of the heat protection of a
full vehicle,” SAE Technical Paper, 2012.
[201] A. Rodionov, F. Mannerhagen, and Y. Liu, “Design and parameter sensitivity
analysis of a heatsink for a direct cooled power module with a ribbon
bonded cooling structure,” in Proc. 2019 21st European Conference on
Power Electronics and Applications (EPE ’19 ECCE Europe), Sep. 2019, pp.
P.1–P.10.
[202] K. Ma, M. Liserre, F. Blaabjerg, and T. Kerekes, “Thermal loading and
lifetime estimation for power device considering mission profiles in wind

193
References

power converter,” IEEE Trans.Power Electronics, vol. 30, no. 2, pp. 590–602,
Feb 2015.
[203] A. Uhlemann and A. Herbrandt, “A new base plate concept on the basis
of aluminium–copper clad materials,” in Proc. PCIM Europe, 2012, pp.
680–685.
[204] K. K. Sikka, K. E. Torrance, C. U. Scholler, and P. I. Salanova, “Heat sinks
with fluted and wavy plate fins in natural and low-velocity forced convection,”
IEEE Trans. Components and Packaging Technologies, vol. 25, no. 2, pp.
283–292, June 2002.
[205] A. I. Zografos and J. E. Sunderland, “Natural convection from pin fin arrays,”
Experimental Thermal and Fluid Science, vol. 3, no. 4, pp. 440–449, 1990.
[206] “OpenFOAM, [Online],” 2020, accessed 2020-11-23. Available at https://
openfoam.org/.
[207] R. Remsburg, “Non-linear fin patterns in cold plates for liquid cooling,” in
Proc. Power Electronics Technology Conference, Long Beach, California,
USA, vol. 10, 2006.
[208] T. Hitachi, H. Gohara, and F. Nagaune, “Direct liquid cooling IGBT module
for automotive applications,” Fuji Electric Review, vol. 58, no. 2, pp. 55–59,
2012.
[209] A. Wintrich, “Thermal resistance of IGBT modules-specification and mod-
elling,” Semikon Application Note, AN1404, 2014.
[210] K. Numakura, K. Emori, Y. Yoshino, Y. Hayami, and T. Hayashi, “Direct-
cooled power module with a thick Cu heat spreader featuring a stress-
suppressed structure for EV/HEV inverters,” in Proc. 2016 IEEE Energy
Conversion Congress and Exposition (ECCE), Sep. 2016, pp. 1–5.
[211] F. P. Incropera, A. S. Lavine, T. L. Bergman, and D. P. DeWitt, Funda-
mentals of heat and mass transfer. Wiley, 2007.
[212] J. Zhu, H. Kim, H. Chen, R. Erickson, and D. Maksimović, “High efficiency
SiC traction inverter for electric vehicle applications,” in Proc. IEEE Applied
Power Electronics Conference and Exposition (APEC), 2018, pp. 1428–1433.
[213] D. Han, Y. Li, and B. Sarlioglu, “Analysis of SiC based power electronic
inverters for high speed machines,” in Proc. IEEE Applied Power Electronics
Conference and Exposition (APEC), 2015, pp. 304–310.

194
Appendix A

The following C-code is used for generating the matrices of the SVM algorithm in
(4.25) and (4.26). Equal distribution of the duty cycles for the redundant states
d1 = d2 (here: duty[0] and duty[3]) is shown in this example, based on 4.24.
The variables from the following C-code are defined as N _layer = N − 1 and
the vectors V _D_abc, V _E_abc, V _F _abc and V _G_abc are calculated from
(4.13):
1 if ( ( u_ref_a_SVPWM + u_ref_b_SVPWM) >
2 ( f l o o r (u_ref_a_SVPWM) + f l o o r (u_ref_b_SVPWM) + 1 ) )
3 {
4 // upper t r i a n g l e sequence : | E F G E | E G F E |
5 duty_E = V_G_ab [ 1 ] − u_ref_b_SVPWM ;
6 duty_F = V_G_ab [ 0 ] − u_ref_a_SVPWM ;
7 duty_G = 1 − duty_E − duty_F ;
8 duty [ 0 ] = duty_E / 2 ;
9 duty [ 1 ] = duty_F ;
10 duty [ 2 ] = duty_G ;
11 duty [ 3 ] = duty_E / 2 ;
12

13 // [ V_E_abc ; V_F_abc ; V_G_abc ; V_E_abc+[1 1 1 ] ] ;


14 v e c t o r [ 0 ] [ 0 ] = V_E_abc [ 0 ] ;
15 v e c t o r [ 0 ] [ 1 ] = V_E_abc [ 1 ] ;
16 v e c t o r [ 0 ] [ 2 ] = V_E_abc [ 2 ] ;
17

18 v e c t o r [ 1 ] [ 0 ] = V_F_abc [ 0 ] ;
19 v e c t o r [ 1 ] [ 1 ] = V_F_abc [ 1 ] ;
20 v e c t o r [ 1 ] [ 2 ] = V_F_abc [ 2 ] ;
21

22 v e c t o r [ 2 ] [ 0 ] = V_G_abc [ 0 ] ;
23 v e c t o r [ 2 ] [ 1 ] = V_G_abc [ 1 ] ;
24 v e c t o r [ 2 ] [ 2 ] = V_G_abc [ 2 ] ;
25

26 v e c t o r [ 3 ] [ 0 ] = V_E_abc [ 0 ] + 1 ;

195
Appendix A

27 v e c t o r [ 3 ] [ 1 ] = V_E_abc [ 1 ] + 1 ;
28 v e c t o r [ 3 ] [ 2 ] = V_E_abc [ 2 ] + 1 ;
29

30 }
31 else
32 {
33 // l o w e r t r i a n g l e sequence : | D E F D | D F E D |
34 duty_E = u_ref_a_SVPWM − V_D_ab [ 0 ] ;
35 duty_F = u_ref_b_SVPWM − V_D_ab [ 1 ] ;
36 duty_D = 1 − duty_E − duty_F ;
37 duty [ 0 ] = duty_D / 2 ;
38 duty [ 1 ] = duty_E ;
39 duty [ 2 ] = duty_F ;
40 duty [ 3 ] = duty_D / 2 ;
41

42 // [ V_D_abc ; V_E_abc ; V_F_abc ; V_D_abc+[1 1 1 ] ] ;


43 v e c t o r [ 0 ] [ 0 ] = V_D_abc [ 0 ] ;
44 v e c t o r [ 0 ] [ 1 ] = V_D_abc [ 1 ] ;
45 v e c t o r [ 0 ] [ 2 ] = V_D_abc [ 2 ] ;
46

47 v e c t o r [ 1 ] [ 0 ] = V_E_abc [ 0 ] ;
48 v e c t o r [ 1 ] [ 1 ] = V_E_abc [ 1 ] ;
49 v e c t o r [ 1 ] [ 2 ] = V_E_abc [ 2 ] ;
50

51 v e c t o r [ 2 ] [ 0 ] = V_F_abc [ 0 ] ;
52 v e c t o r [ 2 ] [ 1 ] = V_F_abc [ 1 ] ;
53 v e c t o r [ 2 ] [ 2 ] = V_F_abc [ 2 ] ;
54

55 v e c t o r [ 3 ] [ 0 ] = V_D_abc [ 0 ] + 1 ;
56 v e c t o r [ 3 ] [ 1 ] = V_D_abc [ 1 ] + 1 ;
57 v e c t o r [ 3 ] [ 2 ] = V_D_abc [ 2 ] + 1 ;
58 }
59

60 // duty c y c l e o f each phase and each s t a t e


61 // A B C
62 // 1 duty_phA_state1 duty_phB_state1 duty_phC_state1
63 // 2 duty_phA_state2 duty_phB_state2 duty_phC_state2
64 // 3 duty_phA_state3 duty_phB_state3 duty_phC_state3
65 // ...
66 // N_layer duty_phA_stateN duty_phB_stateN duty_phC_stateN
67

68 f o r ( i = 0 ; i <=2; i ++) // 3 p h a s e s
69 {
70 f o r ( j = 0 ; j <=3; j ++) // 4 time i n t e r v a l s
71 {
72 duty_state [ j ] [ i ] = 0 ;
73 }
74 }

196
Appendix A

75

76 f o r ( i = 0 ; i <=2; i ++) // 3 p h a s e s
77 {
78 f o r ( j = 0 ; j <=3; j ++) // 4 time i n t e r v a l s
79 {
80 f o r ( k = 0 ; k<N_layer ; k++) // c h e c k t h e s t a t e
81 {
82 i f ( v e c t o r [ j ] [ i ] == k+1)
83 {
84 d u t y _ s t a t e [ k ] [ i ] = d u t y _ s t a t e [ k ] [ i ] + duty [ j ] ;
85 break ;
86 }
87 }
88 }
89 }
90

91 // duty c y c l e o f each s w i t c h
92 // sw1 i s t h e top s w i t c h o f t h e upper−l e g . . . sw_N i s t h e bottom
s w i t c h o f t h e upper−l e g
93 // A B C
94 // sw1 duty_phA_sw1 duty_phB_sw1 duty_phC_sw1
95 // sw2 duty_phA_sw2 duty_phB_sw2 duty_phC_sw2
96 // sw3 duty_phA_sw3 duty_phB_sw3 duty_phC_sw3
97 // . . .
98 // swN duty_phA_swN duty_phB_swN duty_phC_swN
99

100 // sum up t h e duty c y c l e s f o r each s w i t c h


101

102 f o r ( i =0; i <3; i ++) // 3 phase


103 {
104 f o r ( j =0; j <N_layer ; j ++)
105 // s w i t c h i n d e x . " j = 0" c o r r e s p o n d s t o top s w i t c h .
106 {
107 duty_sw [ j ] [ i ] = 0 ;
108 f o r ( k =0; k<=j ; k++)
109 // sum up t h e duty c y c l e from s t a t e N_layer u n t i l s t a t e 1
110 {
111 duty_sw [ j ] [ i ] += d u t y _ s t a t e [ N_layer−1−k ] [ i ] ;
112 }
113 }
114 }

The matrix duty_sw is the final output of the modulator and contains the duty
cycle of each switch of the 3-phase 5-level NPC converter.

197
Appendix A

198
Appendix B

PCBs of Medium Voltage Drive System with NPC Convert-


ers
The PCB layout of the basic boards of the medium voltage generator drive
described in Chapter 5 are attached here.
• In Fig. 1 the PCB that receives the current, voltage and encoder measurement
signals from the measurement devices of the drive is shown. This PCB has
got four copper layers, with the top and bottom layers being available for
connecting the components on the board and the middle two layers being
occupied with the VCC = 3V and the ground potential, respectively. A
picture of this board is shown in Fig. 5.12.
• The PCB in Fig. 2 converts the electrical PWM signal from the microcon-
troller into optical signal through fiber optic transmitters. It is also made of
four copper layers and its schematic has been shown in Fig. 5.13.
• The PCB layout of the gate driver described in Subsection 5.2.1 is shown in
Fig. 3. This board, like all the power PCBs of the system, is made of two
copper layers.

199
Appendix B

Fig. 1: PCB layout of the measurements and hardware protection board: red and blue
lines show the traces of the top and bottom copper layers, respectively

200
Appendix B

Fig. 2: PCB layout of the PWM board: red and blue lines show the traces of the top
and bottom copper layers, respectively

201
Appendix B

(a)

(b)

Fig. 3: PCB layout of the Gate driver for the SCH2080KE SiC MOSFET: (a) top and
(b) bottom copper layer

202
Appendix B

Current Sensor for SiC High-Current Inverter


The IMC-Hall sensors from Melexis have been used for current measurement of
the SiC inverter presented in Chapters 6-7. This type of sensors provides a low
cost and compact solution for current sensing in power converters. The variant
with product code MLX91208-CAV has a large enough measuring range up to
700 A (peak), which is suitable for the specific prototype inverter.
A current measurement PCB has been designed, as seen in Fig. 4(a). The PCB
consists of four layers, with the two top layers (red and green) being used for
routing the signal traces and the third (dark blue) and fourth (light blue) being
reserved for the VCC = 5 V and ground potentials, respectively. The CAD drawing
of the current measurement PCB is shown in Fig. 4(b), while the final assembled
sensor board can be seen on the left side of Fig. 7.4(b).

(a) (b)

Fig. 4: (a) Layout of the current measurement PCB having the MLX91208-CAV sensors
and (b) 3-D CAD drawing of the sensor board mounted on the inverter ac
terminals

203

You might also like