Krastev
Krastev
by
IVAN KRASTEV
University of Birmingham
DOCTOR OF PHILOSOPHY
School of Engineering
University of Birmingham
March 2018
UNIVERSITYDF
BIRMINGHAM
vehicles based on a multilevel cascade converter with full-bridge cells. The converter
provides dc-ac power conversion in a single stage, while compensating for the variation
of fuel cell terminal voltage with load power. The proposed converter can replace the
solutions.
The converter numerical and analytical models are derived showing that the converter
can be modelled as a cascaded boost converter and 3-phase inverter. The design
methodology for the energy storage capacitors and power inductors is presented,
showing that inductance is reduced at a quadratic rate with the addition of more sub-
traction drive in a fuel cell tram demonstrates that the proposed converter is a viable
inverter achieves operation from standstill, with full motor torque, to field weakening
with constant power, boosting a 50V dc supply to 200V peak line-to-line voltage.
ACKNOWLEDGMENTS
This project was a big part of my life for several years, and when I started, I really had
no idea what I was getting into. In the beginning everything was nice and relaxed but,
much quicker than I like to admit, the PhD poured over into my personal life. In all that
time my partner Andriana stood by me, listened to me complaining about stuff blowing
up and was extremely understanding when I’d spend the whole weekend fixing said stuff.
She kept me clean, fed, and most of all - sane. Adi, I truly could not have done this without
you.
I would also like to thank my loving parents for their support and for being so
understanding – I hope I didn’t worry them too much when I was locked away for days in
the basement lab. They also provided priceless advice when it was time to go back to the
real world and find a job and helped me prepare for interviews and negotiations.
The most fun side of the project was technical – designing and building converters,
modelling and doing some embedded coding. Solving my technical problems was made
easy by the wonderful people at the Birmingham Centre for Railway Research and
Education (BCRRE), whom I would like to thank: Pietro Tricoli, my supervisor, was always
available for a technical chat, he let me bounce so many ideas off him (most of them not
very clever), and always had something helpful to say; Dr Stuart Hillmansen got me to
build a hydrogen loco which is how I learned most of the engineering I know; Prof Felix
Schmitt who always lent a helping hand in the right moment; Paul Weston and Louis Saade
were of great help when I started writing code and were invaluable when I had to build
two dozens of PCBs in the space of two months; Prof Clive Roberts for funding some power
electronics shenanigans; and Adnan Zentani, for all the cutting, drilling and machining!
I would also like to thank Nadeen Taylor, for being the best administrator that knows
everything and everyone and is always ready to help out!
Also, a big thanks to all my friends from BCRRE– Heather, Rana, Rob, Krish, Rory, Louis,
Paul, and Rhys. Hanging out, playing board games and watching bad B-movies – those
were lots of fun times and great memories!
Lastly, I would like to express my gratitude to the Engineering and Physical Sciences
Research Council for funding this project, and ©Mitsubishi Electric Corporation for
providing IGBT modules.
TABLE OF CONTENTS
Chapter 2. Review of boost converters and inverter drives for electric traction ..... 9
(IMPBCI) 33
i
2.4. Single stage buck-boost inverters ............................................................................... 40
ii
4.2.2. Average leg sub-module voltage control .......................................................... 91
4.2.4. Injected balancing power for low frequency compensation .................. 111
Chapter 5. General design for an n-module BMCI and experimental converter .. 117
5.1.1. Choice of number of sub-modules n and inverter waveform levels ... 119
6.3.3. Comparison of total inductor energy and inductor losses ...................... 182
7.1.1. Initial validation of the basic operations of the converter ...................... 200
iv
7.2.4. Low-frequency operations .................................................................................. 214
7.3. Experiments with converter driving a simulated traction load .................... 216
8.1.4. Link between passive elements and motor parameters .......................... 230
8.2.4. Optimum carrier modulation to reduce the leg inductance ................... 232
8.2.7. Effects of ddc,x on device utilisation and ac voltage waveform ............ 233
8.2.9. Hybrid traction drive using embedded energy storage ........................... 234
v
Bibliography .......................................................................................................................................... 236
vi
LIST OF FIGURES
Figure 2-1 Diagram of a 2-stage conversion process - dc-dc and dc-ac, with
Figure 2-2 Basic diagram of a traction converter with optional energy storage .......... 10
Figure 2-10 Cascaded H-bridge inverters with separate isolated power sources (left)
Figure 2-15 Multi-phase Interleaved converter with cascaded coupled inductors ..... 34
vii
Figure 2-22 Current-fed quasi Z-source inverter ..................................................................... 44
Figure 3-1 Half-bridge and Full-bridge sub-modules with their respective modulation
Figure 3-3 Schematic diagram for the current flow in MCCs ............................................... 54
Figure 3-5 Structure of a general MCC and difference between full and half bridge sub-
Figure 3-7 Generic sub-module and equivalent circuits used for modelling ................. 59
Figure 3-11 Average-time model block diagram with capacitor voltage and dc-side
Figure 4-1 Staircase modulation for a BMCI with 3 sub-modules and ddc=0.25 .......... 76
............................................................................................................................................................................. 77
Figure 4-3 BMCI switching states for a single leg for two neighbouring voltage steps
............................................................................................................................................................................. 77
Figure 4-4 Half-bridge and full-bridge modules, their PWM modulation signals, and
Figure 4-5 Two modulation schemes that produce the same output voltage ............... 79
Figure 4-6 Modulation and carrier signals, with resultant arm voltages ........................ 80
viii
Figure 4-7 Phase shifted carrier and equivalent APOD carriers ......................................... 81
Figure 4-11 Simplified bode plot of current transfer function, controller, and combined
model ................................................................................................................................................................. 94
Figure 4-12 Simplified bode plot of voltage closed loop, open loop, and compensator
Figure 4-19 Effect of arm duty cycle saturation (above) on filtered arm voltage (below)
........................................................................................................................................................................... 112
Figure 5-2 Modulation for top and bottom converter arms, Phase shifted carrier a.),
Figure 5-3 PWM modulation using APOD carriers a.) and PDC carriers b.) ................ 120
Figure 5-4 Pulsed waveforms of arm and phase voltages, with carrier phase shift of
Figure 5-5 Arm and phase voltage waveforms with 0.5/n < ddc,x < 1/n, and ϑcarrier =
ix
Figure 5-6 THD results at maximum modulation index, for PSC, ϑcarrier= 0°, PDC, ϑcarrier=
Figure 5-7 Instantaneous arm voltages and resultant sum for ϑcarrier = 180° a.) and
Figure 5-8 Arm and dc voltage waveforms with 0.5/n < ddc,x < 1/n, and ϑcarrier = 180°
Figure 5-9 Construction of arm duty cycle trajectory for APOD (left) and PCD(right)
modulation.................................................................................................................................................... 129
Figure 5-10 Duty cycle trajectories for PSC and APOD modulation ................................ 130
Figure 5-11 Duty cycle trajectories for PDC modulation ..................................................... 131
Figure 5-12 Circulating current spectrum for PSC and PDC modulation, ϑcarrier = 0°133
Figure 5-13 Circulating current spectrum for PSC and PDC modulation, ϑcarrier = 180°
........................................................................................................................................................................... 134
Figure 5-14 Circuit and core flux for two separate leg inductors (left) and a single
Figure 5-15 Sub-module arm energy fluctuation for experimental prototype converter,
Figure 5-18 Single converter sub-module showing power PCB with IGBT power
module and control card with microcontroller. The MOSFET leg is not mounted on the
Figure 5-20 Isolated bipolar gate driver with separately isolated power supply ..... 148
Figure 5-21 Digital isolators a.) and drive of gate driver input b.) .................................. 149
x
Figure 5-22 Difference amplifier circuit with current sensing resistor R17 ............... 150
Figure 5-23 Current sensor design and placement with leg inductor ............................ 151
Figure 5-24 Connections between the power stages and the controller. Dashed
Figure 6-1 A Midland Metro Urbos tram (source: Railwaygazette.com) ...................... 156
Figure 6-5 Simplified equivalent circuit model of a fuel cell stack .................................. 160
Figure 6-6 Simulated fuel cell stack voltage-current characteristics ............................. 162
Figure 6-7 Conventional motor drive (left) and proposed BMCI fuel cell drive (right)
........................................................................................................................................................................... 162
Figure 6-9 Feedback values for the BMCI during a traction cycle ................................... 165
Figure 6-10 Converter voltages and currents at start of acceleration showing low
Figure 6-11 Zoomed in capture at t=0.975s, showing dc current ripple caused by the
low-frequency arm-balancing currents; motor currents only show small spikes from the
Figure 6-12 Arm capacitor voltage, arm currents, and circulating current at low
Figure 6-13 Converter in the constant torque region, with ac-side voltage buck ..... 169
Figure 6-14 BMCI under ac-side voltage boost operation .................................................. 170
Figure 6-15 BMCI terminal voltages and currents at constant output power ............ 171
xi
Figure 6-16 Leg inductor voltage and its frequency spectrum during vehicle
Figure 6-17 Converter currents and voltage at start of braking; ac-side voltage
Figure 6-18 Phase leg voltage and its FFT spectrum during vehicle braking .............. 174
Figure 6-26 Lossy BMCI model for a top arm with IGBT switches. ................................. 187
Figure 6-28 BVSI model including conduction and switching losses ............................. 192
Figure 6-29 BVSI phase-to-phase voltages and line currents ............................................ 193
Figure 7-2 Final setup of the converter prototype with a variac load, induction motor
Figure 7-3 Converter phase voltages, referenced to power supply 0V, and output
Figure 7-4 Arm voltages and currents at ddc=0.3, with maximum current ripple ..... 200
Figure 7-5 Converter phase voltages, referenced to power supply 0V, and output
xii
Figure 7-6 Arm voltages and currents for ddc=0.5 ................................................................. 202
Figure 7-7 Converter phase voltages, referenced to power supply 0V, and output
Figure 7-8 Arm voltages and currents for ddc=0.1 ................................................................ 204
Figure 7-9 Sub-module capacitor voltage, line-to-line voltage, and dc-side voltage for
Figure 7-10 Sub-module capacitor voltage, line-to-line voltage, and dc-side voltage for
Figure 7-11 Sub-module capacitor voltage, line-to-line voltage, and dc-side voltage for
Figure 7-12 Comparison between maximum ac-side voltage boost between hardware
Figure 7-13 Submodule voltages for the top and bottom arm of one converter leg. 210
Figure 7-14Average top and bottom arm capacitor voltages during a transient ...... 211
Figure 7-15 Step response of capacitor voltage reference; average arm capacitor
voltages (top) and dc-side and phase currents (bottom) .......................................................... 212
Figure 7-17 Sub-module voltage and arm currents with approximate fout=2Hz ....... 214
cycle................................................................................................................................................................. 217
Figure 7-19 Arms a and b capacitor voltages and currents during the traction cycle
........................................................................................................................................................................... 218
Figure 7-20 Converter capacitor voltages (top) and arm current (bottom) at start of
Figure 7-21 Converter voltages and currents during motoring ....................................... 220
xiii
Figure 7-22 Converter voltages and currents during dynamic braking ........................ 221
mode................................................................................................................................................................ 222
Figure 7-24 Motor speed and capacitor voltage ripple ........................................................ 223
Figure 7-25 Surface plot of capacitor voltage ripple as a function of power factor and
xiv
LIST OF TABLES
Table 2-19 Comparison table for power converters suitable for fuel cell electric
vehicles ............................................................................................................................................................. 50
Table 3-1 Switching states and average output voltage for a half-bridge sub-module
.............................................................................................................................................................................. 52
Table 3-2 Switching states and average output voltage for a full-bridge sub-module52
xv
Table 5-1 BMCI phase voltage levels with different carrier phase shift ........................ 122
Table 5-2 Converter parameters for evaluation of circulating current spectrum ..... 133
Table 5-4 Main parameters of the components of the prototype MCC .......................... 144
Table 6-7 Comparison between simulated BMCI and BVSI ................................................ 194
Table 7-1 Experimental setup parameters for deriving maximum boost characteristics
........................................................................................................................................................................... 205
xvi
LIST OF ABBREVIATIONS
A – Ampere
FB – Full-Bridge
HB – Half-Bridge
HVDC – High-Voltage DC
kW – kilo-Watt
kV – kilo-Volt
xvii
LSC – level-shifted carrier
MW – Mega Watt
PF – Power Factor
Si – Silicon
V – Volt
xviii
VSI – Voltage Source Inverter
W – Watt
xix
xx
Introduction
Chapter 1. INTRODUCTION
Rail transport has been a vital part of moving passengers and goods for more than a
century. Electric locomotives date back to the mid-19th century, when electric motors
became the standard type of traction for tram cars. In the 1890s London Underground
opened the first electric deep underground railway [1], [2]. At the time all electric trains
required an external power supply, that either used a third rail or an overhead catenary
[3]. This was financially viable for city networks and places where steam locomotives
proved too heavy, but for long distance transport of people and goods an autonomous
locomotives were the main type of vehicle as they were able to use multiple types of fuel
and provided the necessary power. By the middle of the 20th century diesel locomotives
were developed, that can use gearboxes and hydraulic torque converters to keep the
diesel engine within its allowed limit of rotations per minute (RPM) [4], [5], but due to the
flexibility of electric drives, diesel-electric engines became the dominant type [6], [7], [8],
[9].
The role of the conversion equipment has always been to match the voltage and
current waveforms of the power supply to those of the traction motors. What has changed
is the efficiency of the conversion process, and the performance that can be achieved by
the motors. Probably the biggest advance in railway traction has been the introduction of
ac motors (both synchronous and asynchronous), that has been made possible by the
availability of fast semiconductors and digital electronics. Ac motors, and the increased
1
Chapter 1
has made them some of the most vulnerable components in the traction drive [10]. [11],
[12], with factors such as rms current and voltage stresses playing an important role in
To this day, diesel-electric trains are the dominant type, as diesel locomotives and
diesel multiple units do not require special infrastructure other than refuelling stations
[15]. Since modern diesel railway vehicles have a fully electric drive, they can be
hybridised by adding a battery or supercapacitor energy storage bank [16], [17], [18],
[19]. However, diesel engines still emit high amounts of NOx gasses, as well as fine
particles, that can cause serious health conditions. This has made diesel technologies
undesirable in urban areas and there is a strong incentive to develop alternatives that rely
Hydrogen fuel cells are such a source. They use hydrogen and oxygen to generate
directly electrical power at a high efficiency. This efficiency is higher than that of diesel
generators [20], [15]: typical diesel generators operate with peak efficiency of 40%,
whereas hydrogen fuel cells can achieve as high as 55% [21], [22], [23], [24]. The only
moving parts of a fuel cell are the hydrogen supply, purge valves, and any air feed pumps
[25]. Compared to diesel generators, maintenance costs are lower [26], [27], and the only
by-products of the hydrogen-oxygen reaction are water and heat. While the quantities of
water are low, and the water can be recycled to cool and hydrate the fuel cell membrane,
the heat is useful in public transport vehicles, as it can be used for vehicle interior heating
The application of fuel cells to railway vehicles requires a suitable design of the power
conversion systems to ensure the required traction control of the motors and the optimal
usage of the fuel cell stacks. However, traditional power converter topologies for fuel cells
powered trains have significant limitations on power density and cost due to passive
2
Introduction
components like inductors and capacitors. This thesis analyses in details the main
requirements of fuel cell traction drives and proposes an innovative power converter
topology suitable for the best exploitation of the capabilities of hydrogen fuel cells.
(kW/litre), specific power (kW/kg), cost (£/kW), and efficiency (%). These characteristics
are measurable and are used to set goals and targets in technology roadmaps [28], [29],
[30]. Depending on the specific application one or two of these parameters will drive the
others. Thus, the designer’s main task is optimizing the power system by choosing a
power converter topology and the main building blocks – traction motor, type of power
state-of-the art technology can be used for any newly designed converter, the main
problem becomes finding the most appropriate topology for the application.
In the automotive sector the number of vehicles is high, with 180 000 plug-in cars
registered in the UK, and counting. Cost and power density stand out as the main drivers
[31], and average power, across vehicles, is around 100kW [31]. Thus, automotive
applications favour simple two-level power converter topologies, where power density
can be improved by increasing device switching frequency. However, they are heavily
The railway sector, on the other hand, has a low number of powered rolling stock –
less than 14 000 in total [34]. Thus, traction systems are tailored to the requirements of
the train and power levels can vary between 10s of kWs and MWs [35]. From this it can
addition, reliability, availability, maintainability, and safety (RAMS) have high importance
3
Chapter 1
[36], [37], as a single train failure can have a propagating effect on a timetable and a single
sources. For example, hydrogen fuel cells, like diesel generators require a power
converter to maintain a constant voltage for the dc-bus of the traction inverters, and any
additional dc-dc converters that would be needed for battery or supercapacitor energy
storage. Like traction inverters, these dc-dc converters are either 2-level unidirectional
with special topologies can improve the power density by reducing the size of the passive
components. However, these approaches have limitations, which are explored in detail in
Chapter 2 and the magnetic components they use have a profound effect on power density
This thesis discusses the Modular Cascaded Converter and how it can be controlled as
a buck-boost inverter, which can be called a Boost Modular Cascaded Inverter (BMCI).
a mixture between a boost converter and multi-level inverter. Unlike traditional power
converters, power transfer from dc-side to ac-side does not occur over one switching
cycle, but over one cycle of the inverter’s ac-side waveform. This has implications both on
Therefore, the main challenges of the design of the proposed BMCI, discussed in this
thesis, are the reduction in passive component size, the choice of number of
requires the smallest inductance, and the control of the proposed converter. The thesis
also discusses the limitations of the BMCI and what areas require further investigation.
4
Introduction
for the design of the passive components, aimed at optimising the power density of the
converter. The suitability for the traction drive of a light rail vehicle is also explored. In
• Compare the BMCI with the current state-of-the-art of traction drives and dc-dc
converters
• Derive the large and small signal models of the converter that can be used for the
• Understand the requirements of the control systems for the operation of BMCI as
• Examine the peculiarities of BMCI design: present and compare the possible
modulation methods, and how they affect the design of the inductors; define
capacitor design method based on system dc-side and ac-side voltages, as well as
5
Chapter 1
• Demonstrate 4 quadrant operation with voltage buck and boost, while operating
with constant torque at low speeds, and field weakening and constant power at
high speeds
1.3. PUBLICATIONS
The PhD project has produced the following works, published in the proceedings of an
The main research paper, that covers a portion of the work presented in this thesis, is
still being developed, with expected submission by the end of February 2019. The thesis
has also generated additional ideas for applications based on BMCI or other modular
6
Introduction
small-signal models.
Chapter 4 - Control: Illustration of the use of the small-signal model from chapter 3, as
Chapter 5 - Design of BMCI: Detailed procedures for the design of a BMCI, including
choice of modulation scheme, inductor design, capacitor design. This chapter also
using data from real-life tram. Presentation of simulation results during a traction cycle,
with special attention paid to converter current and voltage waveforms and harmonic
distortion performance.
7
Chapter 1
8
Review of boost converters and inverter drives for electric traction
This section starts with a discussion of the general requirements for a traction
converter and draws out the specific functionality needed for hydrogen fuel cell vehicles.
A set of metrics are derived to aid the evaluation of different converters and more
modern power converter topologies that could be used in a fuel cell railway vehicle.
The main task is to convert a dc voltage, that can have a range of steady-state values,
different ways and this chapter discusses the following ones: direct dc to ac conversion,
with voltage step-down (buck); indirect dc-dc with voltage step-up (boost) followed by
dc-ac step-down (buck), illustrated in Figure 2-1; direct dc to ac with voltage buck and
boost.
Figure 2-1 Diagram of a 2-stage conversion process - dc-dc and dc-ac, with intermediate voltage Vlink
requirement for a traction converter is that it must supply the full machine current from
0 up to maximum motor speed. To achieve this, the system designer must ensure that the
dc-side voltage can be inverted to produce sufficiently high ac voltage to overcome the
9
Chapter 2
electric motor’s back electromotive force (back EMF). In addition, a modern traction
converter must be capable of 4-quadrant operation, i.e. regenerating energy from the
electric machine. This energy can either be dissipated with a braking chopper [38],
returned to the dc-side power source, or a third port can be added to connect an energy
storage system (ESS) [39], [29], [40], [41], using batteries, supercapacitors, or a 3-phase
Traction
AC
power Motor
Vdc converter
VESS
Figure 2-2 Basic diagram of a traction converter with optional energy storage
The dc-side of the converter can be connected to a power source, such as a fuel cell
stack [44], battery pack [45], [39], [46], or the catenary of a railway network [47], [48],
[49]. These sources are best suited to constant dc current being drawn from them, and
high harmonic content is undesirable. For example, electric railway vehicles have
catenary filter inductors to attenuate low and high-frequency harmonics that can
interfere with communication and signalling equipment [50], [47], [51], but they add to
the traction system’s size and weight. Another important aspect is the level of the dc-side
voltage of the traction inverter, which is affected by the impedance of the power source.
This is particularly relevant for sources like hydrogen fuel cells that have a significant
internal impedance. In fact, when a train accelerates from a stationary position, the fuel
10
Review of boost converters and inverter drives for electric traction
cell voltage is at its peak value, but at full power it is at its lowest. During braking, if the
system uses an energy storage device or a braking chopper, the fuel cell current must be
reduced to 0, which means its voltage will be at its maximum again. Thus, it is beneficial
for a fuel cell drive converter to maintain constant ac-side voltage with varying dc-source
voltage.
The waveform at the ac-side of a traction converter not only controls the electric
machine, but also affects drivetrain reliability. Fast transients at the output of the inverter
can cause reflections and high voltage spikes at the motor terminals [52], [53], while
leakage currents can cause damage to bearings [53], [54], [55], [56], which is a safety
critical part in a railway vehicle [57]. To address these shortcomings the system designer
can use a multilevel inverter, where the phase voltage has 3 or more discrete voltage
levels. This type of converter reduces the peak voltage step and the unwanted effects
An additional benefit is that multilevel inverters cause lower stator current THD,
which in turn reduces the flux oscillations inside the motor. These oscillations cause
losses in the motor yoke and rotor laminations [59], [60], [61]. Multilevel inverters
produce phase voltages that are closer to a sinusoidal waveform and have been shown to
historical problem with railway drives is the limitation of the semiconductor voltage
rating [49], [64]. To balance the high voltage, traction inverters had to be connected in
series, with additional circuitry to equally share the voltage [65]. While current rating can
voltage rating requires complicated gate drive and device protection. At the same time,
using a higher system voltage is the preferred method for increasing the power rating, as
11
Chapter 2
higher current raises conduction losses at a quadratic rate. Thus, topologies that allow for
As discussed in Chapter 1, volumetric and mass power densities are important factors
for traction converters. The biggest contributors to this are the energy storage elements
– the reservoir capacitor and the power inductor [66], [67], [68]. In conventional 2-level
particular, have very poor energy density and specific energy, so alternative approaches
to reducing their size and weight have been the focus of a lot of work [70], [66], [71], [72]
[73], [74], [68], [75]. To judge the suitability of different topologies for traction
converters, the topics discussed above are summarised as the following criteria:
• Energy storage
• Scalability
• Availability
• Reliability
12
Review of boost converters and inverter drives for electric traction
If the traction system only consists of a 3-phase inverter driving one or more ac
motors, the system can manage these conditions with careful design and accepting several
trade-offs. Inverter transistors are selected for the peak fuel cell voltage and peak motor
current. The electric motor is designed with insulation rating that can withstand pulses
equal to the open circuit fuel cell voltage, while windings are dimensioned for current at
the fuel cell voltage at point of maximum power. The resultant drive will effectively be
overrated and will be able to practically deliver up to 60% higher output power.
The inverters discussed are separated into two categories, depending on the number
decoupled by a capacitor [32], [76], [77], with a diagram shown in Figure 2-12. Each
converter leg is switching at a switching frequency fsw and produces 2 voltage levels. The
transistor frequency.
Vdc
Two-level voltage source inverters are the standard converter used in motor drives,
with readily available transistor modules with 6 switches. Due to the inverter’s simplicity
and flexibility of manufacture, it has been used in a prototype fuel-cell drive [78].
13
Chapter 2
As the 2-level VSI uses the minimum amount of semiconductor switches, the topology
has no fault tolerance. If a single open-circuit fault occurs, the electric machine can no
longer be fully controlled, as the amount of state vectors greatly reduces, and if a single
short-circuit fault occurs, the inverter control must be stopped immediately to avoid short
circuit of the dc-link. The properties of the 2-level VSI are summarised in Table 2-1:
frequency
technology
Two 6-transistor bridges can be interleaved using 3 inductors, [79], [80], [81]. Each
inductor has a centre tap that is connected to the electric motor. For a 12-transistor
inverter, the inductor does not require an air gap, but the 6-transistor version does [82].
Effectively, the inverter has a 3-level waveform, oscillating at twice the device switching
14
Review of boost converters and inverter drives for electric traction
Lcc
w1
Vdc v1
u1
Lcc
vu1 vu2
M
vu0
w2
v2 Lcc
u0
u2
Each interleaved inverter provides half the machine current. The dc-side capacitor
RMS current is also lowered, and with a higher ripple current frequency, thus reducing
the capacitor size. As the system comprises of essentially two inverters operating in
parallel, the system is tolerant to a single fault, both switch open circuit and short circuit.
The big drawback of the interleaved inverter is the necessity for magnetic
components, that increase power dissipation and reduce power density. Also, the
interleaved VSI requires 3 additional current transducers for every VSI bridge. The
15
Chapter 2
dc-side current ripple: High Pulsed current, but lower RMS compared to
VSI
level inverter
complicated
A 3-phase inverter can also operate as a current source, where a dc current is chopped
into pulse width modulated currents. The transistor bridge consists of 6 reverse blocking
switches, and the output current is filtered by 3 capacitors, connected directly across the
16
Review of boost converters and inverter drives for electric traction
As traction motors are inductive loads, the output capacitance of the inverter can be
relatively small, with ac-side voltage harmonic distortion much lower than that of a 2-
level VSI.
The CSI requires a configuration of the transistor bridge, that produces a short circuit
for the dc-side current source. This condition can be utilized to boost the dc voltage, and
the CSI is effectively a boost-type converter with a minimum voltage gain of 1. This is not
suitable for traction loads, where motor voltage varies from 0 to Vdc, and an additional dc-
dc converter is required for buck operations. Another drawback is that the CSI’s transistor
bridge can conduct current in only one direction, and to achieve power reversal, the dc-
side voltage must be inverted [45]. The properties of the CSI are listed in Table 2-3.
17
Chapter 2
buck converter
Converters with more than 2 levels have lower ac voltage distortion at full modulation
index, which reduces the iron losses in the electric motor. On average they also have lower
switching losses, for the same effective switching frequency, and lower dc-side capacitor
RMS current.
Multilevel converters can help in reducing the total dv/dt and di/dt of the converter
voltages and currents. They can also utilise devices with lower voltage ratings, with most
topology allowing for higher average phase voltage switching frequency. For inverters the
result is improved total harmonic distortion (THD) of the motor voltage and consequently
This type of inverter requires 2 transistors, with anti-parallel diodes, as well as two
discrete diodes for every additional output level [83], [84]. An n-level inverter requires
18
Review of boost converters and inverter drives for electric traction
the n-1 dc-link capacitors. While the number of devices is per phase, the dc-link capacitors
can be shared between phases, which increases the frequency of the capacitor power
fluctuation and reduces its size. This topology has been used in railway applications [64],
as it allows for a higher dc-link voltage. The 3-level NPC inverter circuit is shown in Figure
Vdc
2
w
Vdc v
M
u vu
Vdc
Vdc
2 Vdc
2
Additional levels require more series connected capacitors, and the dc-side voltage
would be shared equally between them. For reliable operation, the NPC inverter requires
measurement of all the dc-link capacitors and 3 floating current sensors. The capacitors
can be balanced by using external circuits [85] or through the control of the common
mode voltage. In practice, this becomes challenging for inverters with 5 or more levels.
In a phase leg, only 2 devices are switching at any given time, with the other two either
constantly on or constantly off. Over one cycle of the motor waveform, the mean switching
frequency will be less than the modulation carrier, but the transistors will see varying
19
Chapter 2
Vdc
3
w
v
Vdc Vdc M
3 u
vu
Vdc
2Vdc
3
Vdc
Vdc 3
3
While each transistor switch is still rated for Vdc/n, each of the clamping diodes will
require a rating of nVdc/(n-1) [86], which will require either single high voltage diodes, or
dynamic voltage sharing of the dc-side voltage, as device voltages are clamped by diodes
The properties of the neutral point clamped inverters are summarised in Table 2-4:
20
Review of boost converters and inverter drives for electric traction
Switch apparent power: Low Extra pair of diodes per level necessary
Machine current THD: Low to 3-level inverter has a step of Vdc/2, decreased
3 levels
circuit fault
with the same number of transistors as an NPC converter, for the same n. The key
difference is that the FCI does not utilise clamping diodes, and instead has phase
21
Chapter 2
Vdc
2 w
Vdc v
M
u vu
Vdc
Vdc
2
The ideal ac-side voltage waveform is almost identical to the NPC inverter, but is
produced by clamping a phase capacitor to a power rail, or to another phase capacitor. All
devices operate with a constant switching frequency fsw,Q, with the effective switching
frequency of each phase equal to fsw,Q(n-1) [89]. Another quality of the FCI is the available
redundant ac-side voltage states – for voltages not equal to the power rails, there are two
switching combinations that can produce the same voltage step. This degree of freedom
The inverter requires 3 current sensors, and a voltage transducer per each flying
capacitor. Figure 2-9 shows a 4-level MFCI. With n>3, unlike the NPC inverter, the phase
capacitors have different voltages. Thus, the inverter requires multiple capacitors with
22
Review of boost converters and inverter drives for electric traction
w
2Vdc Vdc v
Vdc 3 3 M
u
vu
Vdc
2Vdc
3
Vdc
3
As the only way for current to flow through the phase capacitors is through the load,
the MFCI requires a separate balancing algorithm during start-up. Before capacitors are
charged, their voltage is 0, and the transistors will not share Vdc equally. This is one of the
major drawbacks of this topology. The properties of the FCI are summarised in Table 2-5:
23
Chapter 2
Medium
The cascaded H-bridge inverter uses chains of full-bridge cells, forming 3 converter
arms. They are connected in a star configuration [90], [91], [92], [93], and each end is
connected directly to the traction motor. Each sub-module either has an isolated power
source feeding the H-bridge, or an isolated transformer power link to the main power
24
Review of boost converters and inverter drives for electric traction
M
x=u v w
x=u v w
x = u, v, w
Figure 2-10 Cascaded H-bridge inverters with separate isolated power sources (left) and single dc power source
(right)
For an n-level inverter, the number of devices required is 12n resulting in high ac-side
voltage waveform quality [94]. However, if the converter uses n isolated links per phase,
particularly suited to applications where galvanic isolation between the primary power
As each converter arm supplies one phase of the electric machine, it is essentially a
single-phase dc-ac inverter. Thus, the power in each converter arm oscillates at twice the
Another benefit of CHB converters is the fact that to achieve a higher ac-side voltage,
the converter only requires additional sub-modules [95], which in turn further increases
power quality.
The properties of the Cascaded H-bridge inverter are summarised in Table 2-6:
25
Chapter 2
Switch apparent power: High Isolated converter necessary for each sub-
module
modules
Boost
connected in series, as each phase consists of two converter arms. In contrast to the H-
bridge inverter, the MCC has one inductor connected in series with each arm. The MCC
has been shown to be suitable to motor drives in several publications [63] [96], [97], [98],
[97], [99]. To reduce size each pair of phase inductors can be wound on the same core,
reducing inverter output impedance [100], [101], [102]. Schematic of the converter is
26
Review of boost converters and inverter drives for electric traction
AC Machine
Vdc
vSM,x,n vBatt
vSM,x,n
t
single-phase dc-ac converters. The arm power oscillates at the machine electrical
frequency and at twice that frequency, which is a challenge at low motor speeds.
Different types of sub-module switch configurations have been explored [103]. Most
of them become impractical at low voltage levels, as they require additional diodes. The
The MCC buck inverter half-bridge variant is only suitable for dc-ac operation and is
susceptible to short circuit faults on the dc-side. Thus, for an n–level converter, the
The converter with full-bridge sub-modules can block dc-side short circuit faults and
topologies is that they have reduced device count, while retaining dc-fault ride-through
capability.
The main advantage of MCC-type converters is that increasing the number of voltage
levels is relatively straightforward. Unlike NPC inverters, the MCC does not require
separate balancing converter for n>3. Compared to the FCI, all MCC capacitors have the
same voltage rating. The properties of the MCC are summarised in Table 2-7:
Switch apparent power: High Switches carry both dc-side and ac-side
currents
modules
The single-inductor boost converter is the simplest possible topology, using only 2
semiconductor switches and one inductor, and one or two capacitors, as shown in Figure
28
Review of boost converters and inverter drives for electric traction
2-12. The converter top switch can be a diode, for unidirectional power flow, or a
Vlink
Vdc
The semiconductors need to be rated for the full dc-side current and the peak link
voltage. As the higher voltage rating means increased semiconductor thickness, the total
device area needs to increase, which reduces switching speed. The maximum transistor
volume is roughly proportional to the energy stored in it and increases with the square of
the inductor dc-current. The capacitor ripple current will also oscillate at the switching
frequency, with the maximum amplitude occurring at 50% duty cycle. High capacitor RMS
currents have a detrimental effect on component life and result in increased size and
cooling requirements.
This topology has poor fault tolerance as only a top-circuit short circuit or bottom
switch open circuit faults can be tolerated. If either of these faults occur, the converter can
only transfer power from Vdc to Vlink and it will be without any regulation of Vlink.
This converter is popular for traction applications in the automotive sector and is used
for both fuel cell and battery/supercapacitor onboard systems [114], [113], [111], [115].
Strategies have been developed to reduce the high-side link capacitance [116], [117], but
these schemes cause a higher current ripple in the power inductor. Converter properties
29
Chapter 2
Function: Boost
The single inductor boost converter can be interleaved with additional phase legs,
each one having an inductor, but all connected to the same output capacitor [46], [73],
[111], [118], [119]. For a converter with m legs, each leg’s carrier is phase-shifted by
2fcarrier/m radians. Relative to the individual leg currents, this strategy reduces low-side
current ripple and the its ripple frequency increases proportionally with n. A similar effect
is seen by the link capacitor, reducing the RMS capacitor currents by both increasing
frequency and reducing amplitude. A circuit diagram of a multi-leg converter can be seen
in Figure 2-13.
L1
1 v1 v 2 vm
L2 iL1 iL2 iLm
2
Vlink
Lm
Vdc m
30
Review of boost converters and inverter drives for electric traction
At certain operating points, depending on the number of phase legs, the interleaved
converter can achieve 0 ripple in the low-side and high-side currents. These duty cycle
values are discrete, however, appearing (m-1) times over the duty range of 0 to 1. For
example, a 2-leg converter will have a complete ripple cancelation only for d=0.5. While
interleaving reduces harmonics in the passive component currents, the total volume of
the separate inductors of the boost converter is higher than that of a single inductor
When the two transistor legs have opposing states, a circulating current is introduced,
which is differential for each phase, and consequently does not appear at the output. This
current is circulating between the two converter legs and is effectively the part of the
phase current ripple that does not flow through the low-side capacitor but does flow
dc-side current ripple: Very Low Ripple reduces with number of phases
voltage rating
Function: Boost
31
Chapter 2
extra inductor L0 and winding the individual phase inductors on the same core. When two
converter legs are in opposing states, the inductor L0 voltage is Vdc – ½ Vlink, and it
oscillates at twice the transistor leg switching frequency. The circulating current is limited
by the high differential-mode inductance [75], [120], [68], [66]. A third alternative is to
integrate L0 and the close-coupled inductor LCC into a single magnetic component. In that
case, L0 is the parallel combination of the leakage inductances of each winding [121]. A
L0
Lcc
1
v1 v2
0 2
Vlink
v0 iL0
Vdc
The coupled inductor can be designed with an even higher inductance by utilising the
fact that each side is carrying ½ of the current of L0, resulting in cancelation of any
common mode flux. This means that no air gap is required and the close-coupled inductor
can be made very small. The amplitude of the voltage across the single inductor L0 is also
reduced, as the voltage step at v0 is ½ Vlink, and it oscillates at twice the transistor
converter, the inductor can now be almost 4 times smaller. The reduction is not quite 4-
This converter retains the intrinsic redundancy of the interleaved converter. If one
converter leg fails, the close-coupled inductor core can be saturated, and the converter
32
Review of boost converters and inverter drives for electric traction
can still be operated, albeit with a much higher current ripple. Under normal operation dc
current balance is mandatory between the two converter legs, as any difference between
The converter can be extended to 3-phases by adding an extra winding to the close-
coupled inductor [122], [123]. The interleaved boost with coupled inductor converter
occurs
Function: Boost
The coupled inductor and dc inductor can be integrated into a single component, but
it has been shown that using discrete parts results in a smaller size [75]. A different design
of the magnetic cores, where multiple close-coupled inductors are cascaded, is shown in
Figure 2-15.
33
Chapter 2
v1 v2 v1c
T
1c 1
T
L0 2
0 m-1
Vlink
(m-1)c vm-1 vm v(m-1)c v0 iL0
Vdc m
T
The concept can be further extended to more phases, and the dc inductance L0
progressively reduces in size. However, with more than two phases and a single coupled
inductor, the converter loses the intrinsic redundancy typical of interleaved converters. If
one phase does not carry a dc current the phase’s MMF will be very small, effectively
At the same time the design of the coupled inductor is simplified by the lack of an air
gap, and proximity effect losses depend on the circulating current and dc-current ripple,
The multi-phase interleaved boost with coupled inductor converter properties are
34
Review of boost converters and inverter drives for electric traction
Function: Boost
The flying capacitor boost converter requires 2n transistors for an n-level converter
[72]. It uses a total of n-1 capacitors, but they have different voltage ratings. The FCBC
high-side voltage is no longer symmetrically offset from the link power rails and only
connects to the middle of the series transistor chain, as shown in Figure 2-16.
Vlink
2
Vlink
Vdc
The inductor switching frequency is n times higher than the individual transistor
switching frequency fQ, and converter voltage step is Vlink/n. This is similar to the
interleaved converter topologies, with the difference that the FCBC transistors share the
same inductor current, but the required voltage rating is Vlink/n. The converter can be
35
Chapter 2
scaled up by the addition of extra pairs of transistors and flying capacitors. This converter
has shown to give very high reduction of inductor size, as well as efficiency when using
MOSFETs for the power switches [124]. However, it suffers from the same drawbacks as
The properties of the flying capacitor boost converter are summarised in Table 2-12:
Function: Boost
converters in series [125], [126]. The leg inductor can be split in two, with the return
36
Review of boost converters and inverter drives for electric traction
1 Vlink
2
L0
Vdc
2 Vlink
2
The top and bottom converters can be interleaved to reduce low and high-side
capacitor RMS currents. The converter is especially useful for feeding 3-level inverters, as
it provides a regulated centre-split power supply. This topology is best suited to higher
voltage levels, where transistors with the required voltage rating are readily available and
have lower losses. It has been demonstrated that the TLBC’s inductor size can be smaller
than that of a SIBC or an IBC [127] and can achieve higher efficiency. However, the
converter suffers from a high common-mode voltage switching component as the high-
side link voltage oscillates relative to the negative point of the prime power source.
Function: Boost
37
Chapter 2
Isolated converters allow for galvanic isolation and arbitrary voltage ratios between
power source and link voltage. They still require large capacitors at the low and high-side
terminals, as well as at least one inductor. The main challenges for isolated converters are
balancing magnetic component AC losses between the ferrite cores and the windings
[128]. At the same time, isolated converters can easily achieve zero-voltage switching for
all semiconductor switches, drastically reducing switching losses. The main challenge for
transitions (ZCT), under a wide loading range. This type of system was used on a 1MW
shunt locomotive [15], although exact topology details are not published.
current source driving the transformer. The inductors used can have a small ac flux swing
when low-side voltage has low steady-state variation. The converter can be either in a
push-pull or full-bridge configuration [129], [130], [131]. The two-transistor bridge has
simpler electronics and lower conduction losses but complicates the design of the
L0
iin 1:1:2n Llk
3
Vlink
Vdc
1 2 4
38
Review of boost converters and inverter drives for electric traction
splits the low-side current between the two primary switches. Regardless of the
configuration the converter can achieve regulation over a wide dc-side voltage range, and
it also has a continuous input current. To achieve input regulation the transistor bridge
switches must have a variable overlap to achieve this function, which increases the
Switch apparent power: Medium Active switches required for high efficiency
This topology is a mix between the current fed converter and a dual active bridge
converter. Each leg of the low-voltage side bridge doubles as a boost converter, charging
a local capacitor [132], [131]. When the transformer side is included, the converter
becomes a buck-boost type, and the local capacitor clamps the primary side voltage, acting
as an intermediate link.
39
Chapter 2
L0 v1 v2 v12 v34'
L1 1 3
i1
L2 Vlink
i2 1:n iL0
2 4 i1 i2
Vdc iin
This topology has a continuous low-side current, with lower current stresses on the
primary source, but still requires an ac inductor. However, to achieve wide low-side
voltage regulation the transformer primary voltage is no longer a 2-level waveform, but a
The inductor-fed dual active bridge converter properties be found in Table 2-15:
Switch apparent power: Medium Active switches required for high efficiency
driving a traction motor. These converters usually utilise one transistor bridge for both
40
Review of boost converters and inverter drives for electric traction
dc-dc and dc-ac conversion. Power is provided by a voltage source or a current source, i.e.
inductors and two capacitors [133], [134], [135], [136]. At the low-side the ZSI is
connected to a voltage source, via a transistor. This switch is compulsory for traction
drives, as at low machine speeds the ac-side current is higher than the inductor current.
In this state, the ac-side current must flow through the series connection of C1, C2 and the
L1
Qaux
C1 C2
vZout
M iZin iZout
vZin
vZin vZout
Vdc
L2
The ZSI achieves boost operation by inserting an overlap between the top and bottom
switch of the same leg of the 3-phase bridge. The overlap creates a short circuit across the
output of the impedance network, vZout, that charges the inductor from the energy storage
capacitors C1 and C2, and during the rest of the switching cycle the low-side voltage source
recharges them.
To achieve a higher boost ratio, the duration of the overlap condition needs to increase,
which means the ac-side modulation index reduces. In contrast to the conventional 2-level
VSI, in the ZSI the resultant peak of the fundamental phase-to-phase voltage is lower than
the dc-link.
41
Chapter 2
Thus, the ZSI requires a higher voltage rating of the transistors, compared to a boost
converter plus VSI configuration, but it requires one transistor less. The ZSI input current
has a higher RMS value, as the current drawn from the low-side source has a rectangular
waveform.
or 4 times the switching frequency, resulting in a small inductor value. L1 and L2 can be
wound on the same core, with a high coupling factor, to reduce the total inductance value.
However, Qaux and the 6 bridge transistors are also switching at the same frequency as
vZout, increasing switching losses. Properties of this converter are summarised in Table
2-16.
frequency
technology
Function: Buck-
Boost
42
Review of boost converters and inverter drives for electric traction
A bidirectional QZSI uses the same components as the ZSI, but in a different
arrangement. The low-side voltage source is connected directly to the power inductor,
and the auxiliary switch is moved [137], [138], [139]. Compared to the ZSI the inductor
current is smaller, as it flows through the low-side source, and the two capacitors C1 and
C2 have different steady-state voltages. In other respects, the converter operates the same
way as the ZSI, having the same vZout, ac-side currents, and device stresses.
Qaux
C1
L2 L1 iZout
iZin C2
vZout
M iZin iZout
vdc
Vdc vZout
The QZSI can interface additional power sources by connecting them in parallel with
43
Chapter 2
frequency
technology
Function: Buck-
Boost
This type of inverter has the same transistor bridge as a conventional CSI, with the
input connected to an impedance source network, as well as a third inductor that provides
a connection to the low-side voltage source [140], [141]. A schematic of the converter is
L3
p
C2 iL3 + iL1 + iL2 vdc vpn
w
iu
L2
v M
C1
Vdc u
L1
44
Review of boost converters and inverter drives for electric traction
Like the CSI, the current fed QZSI benefits from the low THD of the output waveform,
but still requires two large capacitors for the impedance network. Importantly, however,
the current fed QZSI is a current source inverter that can buck the ac-side voltage, with
only one additional semiconductor. This auxiliary switch must be either a diode or a
reverse blocking transistor, with the former only usable for voltage boost ratio less than
2 [141]. If a reverse blocking transistor is used instead, higher voltage gain can be
achieved, but similar to the other impedance source converters, the modulation index
While the current fed QZSI has been shown to operate in all 4 quadrants, the transition
from positive to negative power flow is very abrupt, which makes the converter difficult
to control. Additionally, the use of the auxiliary switch is not well explored, and most of
the work on this inverter is done using a diode. This configuration can achieve a voltage
gain less than 2, which is a limitation. Converter properties are summarised in Table 2-18:
45
Chapter 2
technology
Function: Buck-
Boost
general have the advantage of simplicity but are limited by semiconductor technology,
motor current harmonic distortion, ac-side voltage step, and the size of the passive
components.
To address these issues, this thesis presents a new application of the modular
the following problems: size of power inductor, high WTHD, and the need for intrinsic
fault tolerance. The converter has the standard MCC structure but is used to boost the dc-
side voltage. The submodules are constructed using transistor H-bridges, and each one
46
Review of boost converters and inverter drives for electric traction
can produce both positive and negative voltages by inverting the voltage of the local
energy storage capacitor. The schematic of the proposed BMCI is shown in Figure 2-23:
SMa,x,1 ia,x
Ca,x,1
SMa,x,n va,x x = u, v, w
Ca,x,n
L
Vdc vLL v
w
M
x=u ix=u
SMa,x,1 L
Cb,x,1
vb,x
SMa,x,n
Cb,x,n
ib,x
The dc-side capacitor is used to absorb the combined inductor current ripple and, like
other dc-dc boost converters, can be omitted. As with the standard MCC each converter
leg has a pair of inductors, and by increasing the number of sub-modules in each arm, the
required inductance reduces at an almost quadratic rate. A key difference of the BMCI is
that for the same sub-module voltage it can produce a higher peak line-to-line voltage.
The first key benefit of the proposed topology is the big reduction in total leg
inductance, and consequently its weight and volume. This is a consequence of the multi-
level nature of the converter, which also makes the ac-side voltage quasi-sinusoidal. As
this project demonstrates in Chapters 6 and 7, the BMCI can produce machine currents
with very low THD and the maximum ac-side voltage step equal to a-sub-module’s
capacitor voltage. In a practical application, these features are expected to have a positive
effect on traction motor efficiency, as well as motor insulation and bearing longevity.
47
Chapter 2
The second major benefit is that the BMCI can be scaled up in power by increasing the
applications, where there is a significant variability of the power rating for different
trains. At the same time, the leg inductance is a function of effective arm switching
frequency and sub-module capacitor voltage. When the BMCI is scaled with more sub-
modules, the effective arm switching frequency increases, and as long as the system
terminal currents remain constant, power inductor value will actually decrease. In
contrast, for any 2-level boost converter the inductor value increases, as both the peak
A combination of a multilevel inverter and multilevel dc-dc converter could have the
same benefits as the proposed BMCI. However, this type of system cannot achieve the
same availability, as a single switch open-circuit fault will cause a converter to become
inoperable, unless redundant switches are included in the system [142], [143], [144]. The
BMCI has intrinsic fault tolerance because sub-modules can be bypassed, as long as two
of its active switches are operational. Multiple fault management strategies have been
The practical design of a BMCI is discussed in section 5.1 with an example system
simulated in Chapter 6, and also compared to a conventional BVSI design. The results of
the design show that the BVSI topology has 22% higher volume of capacitors and
inductors and 370% heavier mass of the energy storage components. The BMCI achieves
these results at the cost of lower efficiency, which would result in more power that needs
to be dissipated. Albeit this is a shortcoming of this topology, it has to be noted that for
traction applications the energy efficiency has secondary importance to the overall weight
and volume, as any additional weight requires more energy to propel the train, nullifying
48
Review of boost converters and inverter drives for electric traction
the benefits of a higher efficiency and, sometimes, even increasing the total energy
consumption.
more sub-modules
Boost
2.6. SUMMARY
This chapter has examined current state-of-the-art of 3-phase inverter drives, dc-dc
boost converters and single-stage buck-boost inverters, suitable for motor drives of
electric trains powered by hydrogen fuel cells. Different converter topologies have been
railway traction application. Non-isolated converters have a lower cost and can achieve
coupled inductors to reduce the total energy stored in the power inductor. This technique
49
Chapter 2
gives the high overall converter power density, but in practice the number of converter
difficult with more phases and converter fault tolerance decreases drastically.
Isolated converters also offer very high efficiency, but require more transistors, a
power inductor and energy storage capacitors, which limit the maximum achievable
power densities. Additionally, the isolation transformer increases system cost, as it must
To overcome the main limitations of the boost inverter topologies, presented in the
literature, this thesis has proposed to use the modular cascaded converter with full
bridges controlled as a boost modular cascaded inverter (BMCI). This topology achieves
both ac-side voltage buck and boost, while inverting a dc voltage, in a single converter
stage. The proposed BMCI benefits from multi-level ac-side voltage waveform, intrinsic
fault tolerance, and mainly – a reduction in the size of the leg inductors. Unlike the
interleaved converters, discussed earlier in this chapter, the inductance reduction is only
limited by the number of sub-modules. Thus, the converter design can be optimised in
terms of power density or system cost with a balance between the number of sub-
modules, the device voltage ratings, the capacitor voltage rating, and the leg inductance.
Moreover, the BMCI can tolerate short circuit faults both at its dc-side and ac-side
terminals, as well as internal faults in one of its sub-modules. For these reasons, the BMCI
is particularly suitable for applications where the low-side dc voltage reduces as power
A comparison between all the converters reviewed can be found in Table 2-20.
50
Review of boost converters and inverter drives for electric traction
Table 2-20 Comparison table for power converters suitable for fuel cell electric vehicles
Converter Energy Switch Current Voltage Current Scalability Availability Reliability Function
VSI Low Low High High High Low Low High Buck
IVSI Medium Low Medium Medium Low Medium Medium Low Buck
CSI High Medium Low Low Low Low Low High Boost
NPC Low Medium High Medium Low Medium Low Low Buck
to Low
FCI Medium Low High Medium Low Medium Low Low Buck
to Low
CHBI High High Medium Low Low High High Low Buck-boost
MCC High High Low Low Low High Medium Low Buck
VFZSI High Medium High High High Low Low High Buck-boost
QZSI High Medium Low High High Low Low High Buck-boost
CFQZSI High Medium Low Low Low Low Low Medium Buck-boost
BMCI High High Low Low Low High High Low Buck-boost
51
Chapter 2
52
Modelling of Boost Multilevel Cascaded Inverters
This chapter examines the structure and the operation of the boost multilevel
cascaded inverter (BMCI). The converter circuit equations are used to construct a
continuous-time model that can be used for numerical simulations, and then the large and
small-signal analytical models are derived for use in control system design. The maximum
full-bridge and a voltage-source energy storage element, i.e. capacitor, battery or another
power converter. The transistor bridges are controlled with pulse-width modulation to
vary the average ac-side voltage. The half bridge (HB) sub-module uses a single duty cycle
d1 0,1 while the full bridge (FB) has one for each transistor leg – d1 0,1 and
da − 1,1. The duty cycles are compared with a sawtooth (in the case of HB modules) and
The allowed switching states and the average voltage inserted by each sub-module can
be found in Table 3-1 and Table 3-2, while the basic sub-module circuits, and associated
waveforms, can be seen in Figure 3-1. It should be noted that for the same sub-module
output switching frequency the FB carrier is at half the frequency of the HB converter.
53
Chapter 3
Table 3-1 Switching states and average output voltage for a half-bridge sub-module
Q1 Q2 mean (vSM )
vSM
0 0 1 d1VC
+VC 1 0
Table 3-2 Switching states and average output voltage for a full-bridge sub-module
Q1 Q2 Q3 Q4 mean (vSM )
vSM
0 0 1 0 1
-VC 0 1 1 0
vSM (t)
VC f(d1) Q1
VC
d1
d1 t
f(d1) Q2 vSM 0
0
t
vSM (t)
Q1 Q3
VC f(d1) f(d3) VC
d1 da
vSM
Q2 d3 da 0
f(d1) f(d3) Q4 t
t
-VC
Figure 3-1 Half-bridge and Full-bridge sub-modules with their respective modulation and output waveforms
In an MCC n sub modules are connected in series with one another, forming a
converter arm, and a single inductor. The inductor provides current filtering and allows
54
Modelling of Boost Multilevel Cascaded Inverters
Arm a
1
in varm
out
in out Arm b
n
in out
In a 3-phase dc-ac MCC a pair of converter arms form a phase leg, that functions as a
single-phase voltage-source inverter. The output phase current is shared equally between
the top and bottom arm, while the dc-side current is common to both, as indicated in
(3-2):
idc ix
iL ,a , x = +
3 2 (3-2)
i i
iL ,b , x = dc − x
3 2
When operated with a balanced load, the dc-side current flows between the dc power
supply and the converter arms, while the ac-side currents are sourced by the converter
55
Chapter 3
iac
2
idc AC
3 Motor
iac
2
To achieve decoupled control of dc and ac currents, each converter arm has two
separate control variables that are used to calculate the individual sub-module duty cycles
– the dc duty cycle ddc and the ac duty cycle dac. Assuming equal voltage VC across all
floating capacitors, the inserted voltages for each phase x= u, v, w, can be found using (3-3)
and (3-4).
d a , x = d dc, x − d ac, x
(3-3)
d b , x = d dc, x + d ac, x
vb , x − va , x (3-6)
vac, x = = d ac, x nVC
2
In (3-5) the top and bottom arm voltages are complementary, and any large
differences in the resultant dc voltage would cause high currents to be drawn from the
56
Modelling of Boost Multilevel Cascaded Inverters
power supply. For a HB converter, many control schemes keep the average voltage level
Vdc
across the capacitors at VC =
n , resulting in the inserted voltages:
The ac-duty cycle dac,x is calculated using the converter modulation index m, ac-side
Since the maximum duty ratio of a transistor is 100%, the arm duty cycle is limited to
d ac, x (t ) + d dc, x (t ) 0 . If the modulation index is set to 1, the peak ac-side voltage becomes
From (3-12) it is clear that the maximum vac,x,peak occurs at ddc,x=0.5. The peak output
line-to-line voltage range of the HB MCC is limited by the dc-supply. While this implies
that the converter is a pure buck-type, this is not exactly the case. An analogy is a Cuk
converter, where energy from the input is stored in a floating capacitor and then released
into the load. One key difference is that the dc-ac energy transfer occurs over one cycle of
the ac waveform instead of one cycle of the switching carrier. If the arm duty cycle can
have a negative value, the phase ac voltage can go beyond the dc voltage rails, producing
57
Chapter 3
Vdc vuv
vu,b
As the ac-side voltage is now dependent on the average capacitor voltage, the
vac, x , peak =
(1− d )V
dc, x (3-11)
dc
2d dc, x
Equation (3-11) shows that as ddc, x→0, vac,x,peak → ∞. This is typical of boost and buck-
boost converters. As such the MCC has a limit on the maximum achievable boost, that
58
Modelling of Boost Multilevel Cascaded Inverters
SMa,x,1 ia,x
Ca,x,1
SMa,x,n va,x
Ca,x,n
L
Vdc vLL v
w
M
x=u ix=u
SMa,x,1 L
Cb,x,1
vb,x
SMa,x,n
Cb,x,n
ib,x
vSM,x,i vSM,x,i
vC,x,i vC,x,i
vC,x,i vSM or
vSM
t Half t
Full
Bridge Bridge
Figure 3-5 Structure of a general MCC and difference between full and half bridge sub-modules, with typical mean
sub-module output voltages.
while the FB sub-module, with its bipolar output, allows the FB MCC to produce both
positive and negative voltages, that give boost functionality, shown in Figure 3-5.
Therefore, the BMCI proposed in this thesis is based on FB (H-bridge) MCC topology.
numerical simulation, the equations that describe the BMCI are derived. The type of
characteristics can be modelled by a series resistor. The resistive qualities are also
59
Chapter 3
favourable at low dc voltages, where the series connection of multiple IGBTs, and their
saturation voltage drops, can cause severe loss of headroom. Switch current capacity can
quadratic manner.
For this section, the circuit diagram of the BMCI contains the transistor, inductor, and
iIN
SMu,a,1 iu,a
Cu,a,1
rC SMu,a,n vu,a
Vdc
2
Cu,a,n
rL
L vu
LphRph
Vdc x=u
vLL
v
w rL
rC SMu,b,1 L
Cu,b,1
Vdc
rC SMu,b,n vu,b 2
Cu,b,n
ixb x = u, v, w
iD D
D rDS
iD
S S
As the converter is driving a balanced passive load, with line inductance Lph and
resistance Rph, the load star-point voltage is at exactly half the input dc voltage. Since each
phase leg is effectively a single-phase inverter, only one pair of top and bottom arms, with
Each sub-module consists of 4 transistor switches and a floating capacitor, each having
a series parasitic resistance, as shown in Figure 3-7. Parasitic inductances and switch
60
Modelling of Boost Multilevel Cascaded Inverters
capacitances are ignored as they have negligible effect on the average converter
operation.
iC,SM,a/b,x,i
rsw rsw
rC
vC,a/b,x,i fa,a/b,x,iiL,a/b,x
C
Figure 3-7 Generic sub-module and equivalent circuits used for modelling
The sub-module with capacitance C has two circuits associated with it – the one seen
at its terminals and the internal capacitor circuit. Basically, the sub-module capacitor is
seen as a switched voltage source from the rest of the converter, while the capacitor sees
the switched arm current. The transistor states are controlled by the sub-module
switching function fa,a/b,x,i, which has an input da,a/b,x,I and operates according to Table 3-2.
61
Chapter 3
f (d a ,a / b, x ,i , t ) ia / b, x (t )dt (3-13)
1
C
vC ,a / b, x ,i (t ) =
n
n
va / b, x (t ) = vC ,a / b, x ,i (t ) + rSM ,a / b, x ,i (t ) ia / b, x (t ) (3-14)
i =1 i =1
The load is modelled as a voltage source with a series inductance and resistance. This
load, if the 3-phase sources are set to 0. The phase inductance is a point of coupling
between each pair of arm currents and must be included in each of the arm equations:
diL ,a , x di n
n
L + L ph x = − rL + rSM ,a , x ,i (t ) iL ,a , x − R phix − d a ,a , x ,i (t )vC ,a , x ,i (t ) +
dt dt i =1 i =1
(3-16)
V
+ dc − vu (t )
2
(3-17)
diL ,b , x dix n
n
L = − rL + rSM ,b , x ,i (t ) iL ,b, x + R phix − d a ,b , x ,i (t )vC ,b, x ,i (t ) +
− L ph
dt dt i =1 i =1
V
+ − dc + vu (t )
2
Using (3-13) through (3-18), the state-space model of the converter can be derived as
follows:
62
Modelling of Boost Multilevel Cascaded Inverters
M x = Ax + Bu (3-19)
y = Cx
i L ,a , x
iL ,b , x
ix
x= n (3-20)
vC ,a , x ,i
i =1
n
vC ,b , x ,i
i =1
n
n
L rSM ,a , x ,i
−
r +
0 Rph f (d a , a , x ,i ) 0
i =1 i =1
n
n
0 − rL + rSM ,b , x ,i − Rph 0 f ( d a ,b , x ,i )
i =1 i =1
A= 0 0 0 0 0 (3-21)
n
− f ( d a , a , x ,i ) 0 0 0 0
i =1
n
0 f ( d a , b , x ,i ) 0 0 0
i =1
Vdc − v x
− Vdc + v x
B= 0 (3-22)
0
0
L + L ph 0 0 0 0
0 L − L ph 0 0 0
M = 1 −1 0 0 0 (3-23)
0 0 0 0 0
0 0 0 0
0
This model is time-variant, as it depends on switching functions. While it cannot be
directly solved analytically, it is well suited for numerical simulation in Matlab Simulink®.
63
Chapter 3
its sub-module capacitors. The accurate design of the control systems requires an
remove the instantaneous effects of the power transistor switching action state-space
averaging can be used. It is a popular method that derives an equivalent linear time-
invariant system and produces a transfer function, with which classical control methods
can be used.
A single leg of an n-module converter phase leg has n+2 energy storage elements, not
including any load inductance. To simplify the model all capacitors, within the same arm,
are lumped into one element, as it can be assumed that all sub-modules have equal
capacitance, duty cycle, and mean capacitor voltages. This is justified as a real converter
would have a control system dedicated to charge equalisation within the arm, and energy
As the capacitors are the main energy storage elements, the dc-side current control
system will be regulating the average sub-module capacitor voltage for each leg, while the
ac-side currents are controlled by the motor vector controller. This setup is equivalent to
a buck-boost converter cascaded with a voltage source inverter, where the dc-side
converter is regulating the average capacitor voltage, and the inverter is fed, effectively,
by a “stiff” dc-link.
Like section 3.1 the analysis of a single converter leg can be sufficient for modelling
the converter. This is also a practical solution as there would be a separate controller
regulating each leg’s average capacitor voltage. An equivalent circuit of a converter leg
64
Modelling of Boost Multilevel Cascaded Inverters
da,a,xiL,a,x rC
vC,a,x vSM,b,x
C
0.5Vdc
iL,a,u
da,a,xnvSM
rL
Vx La,x
Rx Lx ix rL
Lb,x
da,b,xnvSM
iL,b,u
0.5Vdc rC
C vC,b,x vSM,b,x
da,b,xiL,b,x
x = u, v, w
The switching functions are replaced by a continuous voltage equal to the arm’s duty
cycle multiplied by the corresponding average sub-module voltage. The voltage sources
are averaged over one switching cycle of the semiconductor switches. The capacitor
average voltage, however, is taken over one cycle of the ac-side waveform, as this period
marks a complete energy transfer from the dc-side, to the converter’s capacitors, and then
to the ac-side.
The circuit in Figure 3-8 has two current loops and the corresponding equations are:
65
Chapter 3
di di di
= va , x + rLiLa , x + L La , x + Lx La , x − Lb , x − Rx ( iLb , x − iLa , x )
vdc (3-24)
2 dt dt dt
di di di
+ Rx ( iLb, x − iLa , x )
vdc (3-25)
= vb, x + rLiLb, x + L Lb, x − Lx La , x − Lb, x
2 dt dt dt
For proper inverter operation, the ac-side current must be sinusoidal, without any
The ac-side current iac,x will flow through the dc-side voltage source, but the average
power it gives rise to is 0. Any arbitrary current, limited only by the arm impedances, can
flow between the dc-side voltage source and the converter’s equivalent voltage sources.
However only a dc current would yield an average power to be drawn from the dc-side.
Ignoring current and voltage harmonics, that do not result in net power flow between the
dc source and the load, the dc-side current will be idc,x. Using (3-26), the converter and
iac, x (3-27)
iLa , x = idc, x +
2
iac, x (3-28)
iLb, x = idc, x −
2
If the load current is not shared equally between arms a and b, the resultant current
will circulate power forward and backward between the dc source and the converter,
without transferring it to the load resistor. Substituting with (3-27) and (3-28) in the sum
66
Modelling of Boost Multilevel Cascaded Inverters
di di
vdc = va , x + vb, x + 2rL ( iLa , x + iLb, x ) + 2L La , x + Lb, x (3-29)
dt dt
The arm current equations can be rearranged, like (3-5)and (3-6), to get the dc-side
iL ,a , x + iL ,b , x (3-30)
idc, x =
2
The dc source current flows through the outer current loop in Figure 3-8 and can be
didc, x (3-32)
vdc = va, x + vb, x + 2rLidc, x + 2L
dt
As the load current iac,x is the difference between the two arm currents, it can be
The resultant dc and ac voltages, (3-5) and (3-6), are compensated for the sub-module
capacitor parasitic resistance using vrC ,a / b, x,i = rC d a,a / b, x,i iLa / b, x , which affects the
67
Chapter 3
i
va , x = ( d dc , x − d ac , x ) n ( d dc , x − d ac , x ) idc , x + ac , x rC + vCa , x
2
i
+2nrSW idc , x + ac , x
2 (3-34)
i
vb , x = ( d dc , x + d ac , x ) n ( d dc , x + d ac , x ) idc , x − ac , x rC + vCb , x
2
i
+2nrSW idc , x − ac , x
2
( )
i
va , x = ( d dc , x − d ac , x ) n vCa , x + n d dc , x 2 − 2d dc , x d ac , x + d ac , x 2 idc , x + ac , x rC +
2 (3-35)
i
+2nrSW idc , x + ac , x
2
( )
i
vb , x = ( d dc , x + d ac , x ) n vCb , x + n d dc , x 2 + 2d dc , x d ac , x + d ac , x 2 idc , x − ac , x rC +
2 (3-36)
i
+2nrSW idc , x − ac , x
2
Inside the sub-module, the average leg capacitor voltage and its dynamics can be
described by:
68
Modelling of Boost Multilevel Cascaded Inverters
d vCa , x
C = d a ,a , x iLa , x
dt (3-39)
d vCb , x
C = d a ,b , x iLb , x
dt
dvC i (3-41)
C = d dc, xidc, x − d ac, x ac, x
dt 2
Equation (3-41) shows that the capacitor current can be separated into two
components – one that transfers energy from the dc source, and one that delivers energy
to the ac load, as illustrated in Figure 3-9. The ac quantities dac,x and iac,x are converted to
rC rC
C
da,a/b,xiL,a/b,x vC,a/b,x ddc,xidc,x vSM dac,x,rmsiac,x,rms
C vC
2
69
Chapter 3
(3-42)
(( ) )
i
vdc , x = 2n d dc , x vC + d dc , x 2 + dac , x 2 rC + 2rSW idc , x − 2ddc , x dac , x ac , x ,rms rC
2
(3-43)
(( ) )
iac, x ,rms
vac, x = dac, x nvC − n ddc , x 2 + dac , x 2 rC + 2rSW + 2nddc , x dac , xidc ,x rC
2
According to equations (3-42) and (3-43) the inserted dc and ac side voltage sources
can be modelled as a controlled voltage source, with an equivalent series resistance rSM,
equal to:
( )
rSM = d dc, x + d ac, x rC + 2rSW
2 2 (3-44)
It should be noted that there is cross coupling, caused by the capacitor parasitic
resistance, that can be represented as a voltage source in series with the inserted
capacitor voltages:
iac , x ,rms
vCC ,ac = 4nd dc , x d ac , x
2 (3-45)
idc , x
vCC ,dc = 2nd dc , x d ac , x
2
With (3-42) through (3-45) the differential equations of the dc and ac currents are:
The equivalent circuit of the averaged time model can be constructed using equations
70
Modelling of Boost Multilevel Cascaded Inverters
2vCC,ac,x 2vCC,dc,x nr
2rL 2L 2nrSM SM L+2Lph rL
2ddc,xnvC dac,x,rmsnvC
iac,x,rms 2Rx
Vdc idc,x 2
C
ddc,xidc,x vC dac,x,rmsiac,x,rms
2
x = Ax + Bu (3-48)
y = Cx
Where x is the state variable vector, u is the model input, A, B, and C, are matrices, and
idc, x
x = vC
(3-49)
iac, x ,rms
2
u = v IN
The state variable matrix A, and input matrix B can be compiled from (3-46) and (3-47):
r + nrSM nd dc, x 4n
− L − d dc, x d ac, x ,rms rC
L L L
d dc, x d ac, x ,rms (3-50)
A= 0 −
C C
4n nd ac, x ,rms rL + nrSM + 2 Rx
−
L + 2 L d dc, x d ac, x ,rms rC L + 2 Lx L + 2 Lx
x
1
2L (3-51)
B= 0
0
The output matrix C can have multiple forms, depending on the required output:
71
Chapter 3
CVc = (0 1 0) (3-52)
Cidc, x = (1 0 0) (3-53)
The dc current can be calculated using (3-53), the average sub-module capacitor
The large signal model determines the dc operating point of the converter, around
which the model is averaged. For an inverter with a sinusoidal output, this is a steady state
condition for which the amplitude and phase of the output current are constant. The large
signal model is constructed by setting all derivatives to 0. The steady-state state variables
x = A−1Bu (3-55)
y = CA−1Bu
The dc transfer characteristics, i.e. the dc-side voltage to sub-module capacitor voltage
transfer function, can be obtained by solving (3-55) using (3-52). The steady-state RMS
m (1 − d dc , x ) (3-56)
d ac , x ,rms = ,
2
, where m is the steady-state modulation index. To calculate the maximum gain of the
converter for a certain dc duty cycle, m will be considered equal to 1. The transfer
cycling through values of ddc,x. At every iteration (3-56) and (3-50) must be updated.
72
Modelling of Boost Multilevel Cascaded Inverters
The small signal model describes the dynamics of the converter around the steady-
state operating point. The model gives information as to how the converter dc current and
average capacitor voltage, change relative to changes in the control input – the dc duty
cycle ddc,x. This is achieved by introducing a small perturbation around the dc operating
point, derived in the previous section. The “rms” subscript is omitted from small signal
d dc , x = d dc , x + Ddc , x
idc , x = idc , x + I dc , x
vdc = vdc + Vdc
(3-57)
vC = vC + VC
iac , x ,rms iac , x I ac , x ,rms
= +
2 2 2
d ac , x = Dac , x
In (3-56) the ac duty cycle is related to the dc duty cycle and this equation defines dac,x
as a steady-state variable. Substituting with (3-57) in (3-42), (3-43), and (3-41), and
removing any purely dc variables and cross-coupled components, the converter equations
become:
73
Chapter 3
didc , x rL + nrSM n n
=− idc , x − Ddc , x vC − VC d dc , x
dt L L L (3-58)
4nDdc , x Dac , x ,rms rC iac , x 4nDac , x ,rms rC I ac , x ,rms v
+ + d dc , x + dc
L 2 2L 2L
~
dv~C ~ I dc, x Ddc, x ~ Dac, x iac, x (3-59)
= d dc, x + idc, x −
dt C C C 2
Similar to the general model, the state-space matrices can be populated using the 3
differential equations that describe the converter. However, now there are two model
The duty cycle ddc,x would be used for analysing the control-to-output characteristics.
x = Ax + BIN u + Bd d
idc , x = Cidc , x x (3-61)
vC = CVc , x x
vRx = CVRx x
74
Modelling of Boost Multilevel Cascaded Inverters
r + nrSM nDdc, x 4n
− L − Ddc, x Dac, x ,rms rC
L L L
Ddc, x Dac, x ,rms (3-62)
A= 0 −
C C
4n nDac, x ,rms r + nrSM + 2 Rx
− L
L + 2 L Ddc, x Dac, x ,rms rC L + 2 Lx L + 2 Lx
x
1
2L (3-64)
BIN = 0
0
CVc ,x = (0 1 0) (3-65)
Cidc, x = (1 0 0) (3-66)
With the small-signal model completed, the converter dynamics can be finally
modelled. To design the gain of a closed-loop controller, the required transfer function
can be generated from (3-62)through (3-67), with the block diagram in Figure 3-11:
vC
CVc,x
ddc,x x 1 x
Bd s idc,x
Cidc,x
A
Figure 3-11 Average-time model block diagram with capacitor voltage and dc-side current outputs
75
Chapter 3
Typical control systems for high-power converters use nested control loops, with an
inner current and an outer voltage loop. This controller configuration allows for average
current limiting and to be implemented, the required transfer functions are ddc,x to idc,x and
ddc,x to vC:
idc, x (s )
~
= Cidc, x (sI − A) Bd u
−1 (3-68)
d dc, x (s )
~
v~C , x (s )
= CVc , x (sI − A) Bd u
−1
(3-69)
d dc, x (s )
~
3.4. SUMMARY
The goal of this chapter was to present the basic operating principle of the BMCI and
elaborate on that with detailed modelling of the converter. The equations, that describe
the continuous-time behaviour are derived to aid numerical simulation of the converter.
These are used in Chapter 5 to create a Simulink® model. With the differential equations
of the converter defined, the circuit is simplified to derive the dc transfer characteristics,
which give information on the maximum achievable boost ratio and the effects of parasitic
resistances on the converter performance. The simplified circuit is then used to create the
small-signal model, which is necessary for the design of a stable control system.
An important conclusion is that the while the BMCI appears to be very complex, with
response can be modelled as a 3-rd order system, with only 3 energy storage elements –
inductor L, sub-module capacitor C, and ac inductance L+2Lx. This allows for conventional
76
Modulation and control system of BMCIs
This chapter discusses the basics of BMCI modulation and the design of the converter
control system. The modulation determines the transistor switching pattern and makes
each converter arm appear as controlled voltage source. By calculating the appropriate
voltage, the BMCI controller can maintain the required arm currents to transfer power
The various systems that control the converter are analysed from the lowest level, i.e.
controllers that operate on the converter’s internal variables, up to the highest level –
controllers that make the converter appear as a 2-terminal “black box”. This chapter uses
equations and models described in Chapter 3. A novel controller for sub-module capacitor
voltage balancing is presented in 4.2.1.2, that avoids the need for gain tuning.
A controller design example for a prototype fuel cell BMCI drive can be found in
Appendix B.
4.1. MODULATION
The BMCI’s converter arms are essentially voltage source converters connected in
to limit the converter current and provide voltage boost or buck functions. To apply a
more precise control to the inductor current, different modulation schemes can be
employed.
The staircase method is the simplest modulation technique. It essentially samples the
modulating waveform at discrete intervals, with the resolution depth equal to the number
77
Chapter 4
Figure 4-1 Staircase modulation for a BMCI with 3 sub-modules and ddc=0.25
The offset in the modulated signal, equal to ddc,x, causes the arm voltages to be
asymmetric around the average value. This means in turn that the output phase switching
levels increase. This is valid for converters with an inductive load, such as electric motors.
The increase in the number of voltage levels is made possible by the two inductors in the
converter leg. The difference between the dc-side voltage and the dc-component of the
summed arm voltage is impressed across the two inductors. As this voltage has steps of
This property can be better examined by first equating a full transistor bridge to a 3-
78
Modulation and control system of BMCIs
VSM
Each sub-module can either insert a positive, a negative voltage, or 0V. For achieving
a certain phase voltage, the switching functions are defined by the modulation signal of
each arm, as shown in Figure 4-1. Between two neighbouring levels, the circuit of a single
converter leg is shown in Figure 4-3. As the two arms can be switched independently, the
by a half of it depending on whether the sum of the individual arm voltages is even or odd.
iL,a,x iL,a,x
0.5Vdc VC 0.5Vdc VC
vxa=-1VC VC
vxa=-2VC VC
VC VC
x = u, v, w x = u, v, w
L L
Rx Lx Rx Lx
vL,x=Vdc -2VC vL,x=Vdc -1VC
ix ix
vx=2VC vx=2.5VC
L L
0.5Vdc 0.5Vdc
VC VC
vxb=3VC vxb=3VC
VC VC
VC VC
iL,b,x iL,b,x
Figure 4-3 BMCI switching states for a single leg for two neighbouring voltage steps
Staircase modulation in general requires large inductors to limit the arm currents.
Better performance is achieved with high values of n, where the voltage steps are
sufficiently small. The staircase modulation method is unsuitable for BMCIs because to
achieve stable operation, the dc duty cycle ddc,x must be divisible by 1/n.
79
Chapter 4
With PWM the arm voltages switch between the two closest levels to the modulation
comparator would generate the pulse signals for each transistor half-bridge. The carrier
module, the control duty cycle is directly applied to the top transistor. In a full-bridge (FB)
sub-module, two duty cycles are required, one for each transistor leg and the resultant
sub-module output voltage vSM will be switching at twice the individual leg switching
frequency, fsw,SM = 2fsw,Q. The sub-module leg duty cycles are symmetric around 0.5 and the
difference between them is equal to the active duty cycle da. This is shown in Figure 4-4:
vSM (t)
f(d1) Q1
VC
d
d t
f(d1) Q2 vSM 0
0
t
vSM (t)
Q1 Q3
f(d1) f(d3) VC
d1 da
vSM
Q2 d3 da 0
f(d1) f(d3) Q4 t
t
-VC
Figure 4-4 Half-bridge and full-bridge modules, their PWM modulation signals, and output voltages
For the same sub-module output voltage frequency, the half-bridge carrier must be
oscillating at twice the frequency of the full-bridge carrier. As the control variable of each
sub-module is the active duty cycle, that can be both positive and negative, the individual
80
Modulation and control system of BMCIs
1 + da
d1 =
2 (4-1)
1 − da
d3 =
2
The same voltage output, with duty cycle da can be produced if the sub-module uses
one control signal and two carriers, as shown in Figure 4.5 a).
v(t)x,a/b,i
da
-d3
+Carr ier
Carr ier
t t
d1 -Carr ier
vSM(t)
a.) b.)
Figure 4-5 Two modulation schemes that produce the same output voltage
This modulation method, however, has each half-bridge either switching at twice the
original carrier frequency, or being constantly off. In a real converter the modulation in
Figure 4-5 b.) will result in higher variance in switching and conduction losses over one
full cycle, as one transistor leg will be switching at 2fsw,Q, while in the other leg only one
The various methods for applying PWM to the whole of the converter can be classified
in two general types – phase shifted, and level shifted. In phase shifted carrier schemes,
the converter sub-modules are constantly switching with frequency equal to 2fsw,Q.
Instead, level-shifted carrier schemes assume that only one sub-module’s voltage, within
an arm, is modulated at any given time, and the duty ratio cycles from 0 to 100% n-times
81
Chapter 4
When a phase shifted carrier is used, each sub-module generates its own carrier
waveform. Each sub-module’s carrier is phase shifted by 1/nth of the switching period of
vSM. With phase-shifted carrier modulation, the series connected sub-modules are
interleaved. This results in an effective switching frequency equal to 2nfsw,Q, and similar
to a parallel interleaved converter, as the modulation signal is varied, so does the duty
cycle of the switching portion of the arm voltage. An example is shown in Figure 4-6,
n n
n da,x n
2 2
n n
1 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 2 db,bx
n n
n n
n n
Figure 4-6 Modulation and carrier signals, with resultant arm voltages
The modular cascaded converter’s duty cycle resets back to either the maximum or
the minimum value as the arm’s modulation signal crosses an intersection of the
interleaved arm carriers. Effectively each intersection, at a value of 1/n, defines an arm
voltage level step. Thus, the phase shifted carrier modulation can be equated to a level
shift scheme, the alternate phase opposition disposition carrier. The two waveforms
produce the same arm voltage for all values of ddc,x and dac,x. A comparison between the
82
Modulation and control system of BMCIs
Tsw,Q Tsw,Q
n 2 n 2n
n n
2 2
n n
1 dac 1 dac
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 2
n n
n n
n n
Phase shifted carrier can be implemented with either a centralised carrier generator,
or a localised sub-module carrier generator. The main benefit is that as for │da│<1, all
sub-modules in the converter arm will have the same switching frequency.
In a level-shifted carrier (LSC) scheme, each converter arm level has a corresponding
carrier, oscillating at the same effective arm switching frequency as the PSC schemes –
2nfsw,Q. As the number of active, i.e. non-zero, inserted sub-modules increases, some sub-
modules are in a static state of producing either ±VSM, or 0. For level shifted carrier
For this carrier type to be used, it has to be either generated in a single central
controller, with the transistor signal sent directly to the sub-module transistor gate
drivers, or if each sub-module has its own carrier, it has to be phase shifted by 180° for
values of da/b,x<0.
83
Chapter 4
In this modulation method all carriers have the same phase, for both positive and
n n
n da,x n
2 2
n n
1 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 2 db,bx
n n
n n
n n
PD modulation for BMCI converters maintains the same direction of the pulse edges
throughout the full range of duty cycle values, which will be shown in Chapter 5 to give
This carrier modulation method is similar to PD, with the key difference that for
84
Modulation and control system of BMCIs
n n
n da,x n
2 2
n n
1 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 2 db,bx
n n
n n
n n
With the POD scheme when the modulation signal of one arm enters the negative
region, the alignment between the top and bottom arms changes, giving similar results to
the APOD scheme, which is identical to the conventional PSC modulation. This modulation
scheme has not been examined in detail and can be included in a future review.
The biggest difference between space-vector modulation (SVM) and the previous two
converter simultaneously. This method creates a vector space, with discrete switch states
that correspond to fixed ac-side voltages. This modulation utilises the redundant states,
i.e. different combinations that produce the same ac-side voltage, to reduce switching
SVM does not require the generation of PWM carriers and has intrinsic third harmonic
injection. One drawback of SVM method is that the number of possible states increases
85
Chapter 4
The BMCI can potentially use a small n, such as the experimental prototype described
in this thesis, that has a total of 12 cells. In such a case the complexity of the SVM would
not be very high, and it is certainly a viable method that can be explored in future works.
voltages to a desired value, maintain equal sub-module capacitor voltages across the
otherwise known as Proportional Integral (PI) compensators, and the section covers the
The novel contribution in this section is the voltage feedforward arm balancing
algorithm, described in 4.2.1.2. The benefit of the proposed method is that no gain tuning
A key requirement for any MCC is that all sub-module capacitor voltages are equal.
This condition is necessary to achieve high ac-side voltage quality, and consistent current
capacitance spread of ±20%. In a converter arm all sub-module transistor bridges have
the same arm current flowing through them, and if they all operate with the same duty
cycles, meaning equal capacitor currents, the differences in capacitance will result in
unstable operations of the converter. For converter using a phase-shifted carrier, the
individual sub-module duty cycle will be the same as the arm duty cycle:
86
Modulation and control system of BMCIs
d a , x ,i = d a , x
(4-2)
d b , x ,i = d b , x
d 1 (4-4)
VC ,a / b, x ,i = da / b, x ,i iL ,a /b , x ,i
dt Ca / b, x ,i
For a level-shifted carrier, m sub-modules will have a 100% duty cycle, l sub-modules
will have 0% duty, while 1 will have a duty cycle equal to the local duty:
(
m = floor nd a / b , x )
l = n − m −1 (4-5)
nd a / b , x
d a / b , x ,local = d a / b , x −
n
Unlike 2-level dc-dc converters, where the net power of a capacitor is 0 over one
switching cycle, in an MCC the net power is 0 over one cycle of the output sinusoid. For
this reason, the sub-module power can be considered continuous over one switching
pC ,a /b, x,i = iC ,a /b, x,i vC ,a /b, x,i = da /b, x,i iL,a /b, x vC ,a /b, x,i (4-7)
d 1 (4-8)
VC ,a / b, x ,i = d a / b, x ,i iL ,a / b, x
dt Ca / b, x ,i
The sub-module capacitor voltage vC,a/b,x,i must always be positive, as the converter
uses transistors with anti-parallel diodes, and thus the capacitor power is dependent only
on the arm current and the sub-module duty cycle. The arm duty cycle da/b,x is used to
87
Chapter 4
control the converter input, output, and leg currents, while the arm current is determined
To achieve an equal sub-module capacitor voltage across each converter arm, the only
control variable is the individual sub-module duty cycle. How this is modified, from the
value determined by the higher-level control system, is where different types of balancing
When local feedback is used, the balancing duty cycle da/b,x,i,bal is the output of a
proportional controller. The controller input error is calculated from the average arm
The average arm duty cycle can be calculated for each time period TSW,Q = 1/fsw,Q using:
The biggest drawback of this balancing controller is that the mean value of da/b,x in
(4-10), after the controller has been executed according to equation (4-9), will not always
be equal to the original value of da/b,x. This will cause a disturbance that will affect the
The dynamics of this scheme can be evaluated using the fact that only dbal will cause a
dc change in arm capacitor voltage, as the energy stored in the arms caused by ddc,x and
88
Modulation and control system of BMCIs
(
sCvC ,a / b, x ,i = d a / b, x,rms iL,a / b, x,rms cos ( x ) k p,bal vC ,a / b, x − vC ,a / b, x,i ) (4-11)
1
vC ,a / b , x ,i = k p ,bal vC ,a / b, x
d i cos x (4-12)
C s + k p ,bal a / b , x ,rms L ,a / b , x ,rms
C
This method scales the duty cycle of each converter sub-module based on the
individual sub-module capacitor voltage. When the arm power is negative, i.e. the
capacitor is discharging, the scaling factor of the i-th sub-module duty cycle is equal to the
ratio of i-th capacitor voltage over the mean capacitor voltage for the whole arm:
vC ,a / b, x ,i
d a / b , x ,i = d a ,b / x for pa / b, x 0 (4-13)
vC ,a / b, x
Since the total arm voltage equal to the sum of all sub-module voltages, and the scaling
factors are proportional, the mean duty cycle of the whole arm remains equal to da/b,x over
a complete sub-module switching cycle. At the same time, the module with the highest
voltage provides the highest power, while the one with the smallest charge supplies the
lowest power.
When the arm power is positive, i.e. the capacitors are charging, the ratios need to be
inverted. However, to maintain mean duty cycle equal to the command value, the scaling
factors need to use the difference between twice the mean sub-module capacitor voltage
89
Chapter 4
Similar to the proportional feedback controller, the current that causes a change in the
capacitor voltage is the difference between the sub-module current and the mean current:
v
sCvC ,a / b, x,i = d a / b, x,rms iL ,a ,b, x ,rms cos ( x ) 1 − C ,a / b, x ,i (4-15)
vC ,a / b, x
1
vC ,a / b, x ,i = vC ,a / b, x
vC ,a / b, x (4-16)
sC +1
d a / b, x ,rms iL ,a / b, x ,rms cos x
According to (4-16) the feed-forward balancing controller will also have a first-order
system response, as does the proportional controller. The scaling term for each sub-
module is:
vC ,a / b, x ,i
k a / b , x ,i = (4-17)
vC ,a / b, x
The block diagram of the converter sub-module capacitor balancing control system is
shown in Figure 4-10. The system used in this thesis is the feedforward control, discussed
in section 4.2.1.2.
ka/b,x,i
2
vC,a/b,x,i
vC,a/b,x
ia,x/bda,x sign
90
Modulation and control system of BMCIs
Another method for balancing the capacitor voltages is to use a sorting algorithm. The
general idea is that at every switching instant, the control system evaluates all the sub-
module voltages and chooses which one will produce 0 volts, i.e. neither charging or
discharging, or ±vC, charging or discharging. As with the previous methods discussed, the
rule changes depending on arm power flow, in other words it depends on whether the
For positive power flow into the arm, pa/b,x>0, the next sub-module to be inserted is
the one with the lowest state of charge, and the one to be bypassed, is the one with the
highest state of charge. For negative power flow, or arm discharging, the next module to
be inserted is the one with the highest state of charge, while the next one to be bypassed
Compared to the methods discussed in sections 4.1.2.3 and 4.1.2.4, sorting algorithms
are more computationally intensive for high number of sub-modules. The rate at which
the sorting algorithm is executed, fSort, can be either the same as 2fsw,Q, or lower. Making
fSort much lower than fsw,Q will reduce the effectiveness of the balancing control system,
and in addition fSort must be higher than the converter output frequency, fOUT=2πωe, in
order to track the arm power polarity. The sub-module capacitance determines how quick
the balancing action is, and thus higher capacity can use a lower fSort. As the module being
modulated is the one with the highest or lowest charge, this method cannot guarantee
The goal of the balancing control systems in 4.2.1 and 4.2.2.3 is to only equalise the
voltage across sub-modules in the same converter arm, converging to the mean arm
voltages. When the converter is operating without a closed loop controller, the actual
91
Chapter 4
capacitor voltage and output currents are determined by the modulation index and dc
duty cycle ddc,x, as well as the load resistance and converter parasitic elements.
Similar to 2-level dc-dc converters, the BMCI can use an outer voltage control loop and
an inner current loop [73], [152]. This configuration allows for a more reliable
conventional dc-dc converters, the BMCI has additional controllers to manage any voltage
differences between arms a and b within the same arm, suppression of unwanted
controller is necessary in all cases, and its current reference is derived from the outer
voltage controllers.
The circulating current is common to both arms a and b and flows through the dc link
of the converter, and controls the input power for the converter leg of phase x:
n
pdc , x = d dc , x ( vC ,a , x ,i + vC ,b, x ,i ) idc , x (4-19)
i =1
In chapter 3 the mathematical large-signal and small-signal models are derived for a
single converter leg. The control input to circulating current transfer function can be
derived by converting the state-space model from equations (3-62) through (3-64), using
the output matrix for the state variable idc,x from (3-66):
92
Modulation and control system of BMCIs
x = Ax + Bd d dc , x (4-20)
y = Cidc x
As the detailed model includes the dynamics of the ac circuit, the model is of the 3rd
order, as it has 3 energy storage elements. This complicates the analysis and the design of
the control systems. The phase inductance Lx and the load resistance introduce an
additional pole in the system response. If that pole frequency is much higher than the
natural frequency of the plant ωn, the effect of the phase inductance becomes insignificant
to the input current controller and can be removed from the state-space model. The phase
resistive load can be replaced by equivalent sub-module resistive load Rx,SM. This
Rx
2 + rL + nrSM (4-21)
cos
Rx ,SM =
nDac , x , RMS 2
The simplified equivalent circuit leads to revised matrices A2, Bd2, and Cidc:
rL + nrSM nDdc , x
− −
L L
A2 = (4-22)
Ddc , x 1
C CRx , SM
n
− L VC
Bd ,2 = (4-23)
I dc , x
C
Cidc = (1 0 )
(4-24)
Cvc = ( 0 1)
The converter’s dc duty cycle, ddc,x, to circulating current, idc,x, transfer function can be
93
Chapter 4
idc , x ( s )
= Cidc ( sI − A2 ) Bd 2
−1
H idc ( s) = (4-25)
d dc, x ( s )
s
+1
idc , x ( s ) z (4-26)
H idc ( s) = = Gidc
d dc , x ( s ) s
2
2 s
+ +1
n 2
n
, where Gidc is the dc current gain, ωz is the plant zero, ωn is the plant natural frequency,
and ζ is the damping ratio. With the revised matrices, the model order is reduced to
second, and the equations are identical to those of a conventional boost dc-dc converter.
The equivalent zero frequency, natural frequency, and dc current gain for the BMCI are:
2 (4-27)
z =
RSM C
2nVC
Gidc , x ( 0 ) = − (4-29)
nd RSM + rarm
2
converter, stability over a wide range of loads, and converter response. To achieve
stability, the open loop phase shift at the crossover frequency must be less than 180°, and
a phase margin of more than 60° will avoid excessive setpoint overshoot. To maintain
stability over a range of loads, and changes in leg inductance, a gain margin of at least 6dB
is required.
94
Modulation and control system of BMCIs
As the control voltage of the BMCI is averaged over one cycle, of the effective switching
frequency fsw,eff=2nfsw,Q, the arm duty cycle will be updated once per effective switching
cycle, resulting in a discrete system. This effect can be modelled using a Zero Order Hold
s
−
1− e fsw,eff (4-30)
H ZOH = f sw,eff
s
At the desired crossover frequency, equal to 10% of the open loop crossover
frequency, the ZOH introduced 18° to the phase response, which is equivalent to a delay
of one-half switching cycle. The 0dB crossover frequency of the open-loop current plant
idcGidcn 2 (4-31)
i ,0 dB =
z
Since the proportional gain coefficient, kp, sets the crossover frequency of the cascaded
i ,CL (4-32)
k p ,idc =
i ,0 dB
The integrator gain, ki,idc, is used to set the controller zero. Above the natural
frequency, the magnitude response of the plant decreases with frequency at a rate of
20dB/decade, meaning the phase shift is 90°. Below the compensator zero frequency,
ωz,comp the phase shift is also 90°, while at the break frequency it reduces to 45°. The
compensator zero ωz,comp is set to one decade below the crossover frequency, so that it
95
Chapter 4
The open loop transfer function can be evaluated for crossover frequency, phase, and
gain margin using the forward transfer function of the plant and compensator:
The simplified bode responses of the different transfer functions is shown in Figure
4-11:
i,0dB,CL
z n i ,0 dB Frequency (rad / s)
Phase ( o)
-90o
Frequency (rad / s)
o
0
90o
180o
Figure 4-11 Simplified bode plot of current transfer function, controller, and combined model
The stability of the system can be evaluated by analysing the Nyquist plot of the
cascaded open-loop – compensator transfer function. The closed loop gain of a stable
96
Modulation and control system of BMCIs
, where iref,x is the reference current variable. The compensator introduces additional gain
below the natural frequency of the converter current transfer function, removing steady-
state error. Thus, the transfer function of the closed loop current system can be
1
H idc ,CL (4-37)
s
+1
0 dB ,CL
The BMCI can operate with voltage mode control, with the control input being ddc,x.
The transfer function of duty cycle to capacitor voltage transfer function can be extracted
s
−1
vC RHP (4-38)
H vc = = Gvc
2 s
2
d dc , x s
+ +1
2
n n
Gvc = −
(
I dc , x RSM nDdc , x 2 RSM − rarm ) (4-39)
rarm + nDdc , x RSM 2
derived. To design the capacitor voltage control system, the required transfer function is
reference current iref,x to sub-module capacitor voltage vC,x. This can be found using
97
Chapter 4
vC , x d dc , x idc , x H vc (4-40)
H vc ,iref = = H idc ,CL
d dc , x idc , x iref , x H idc
s
Gvc − 1
H vc v
= C = RHP (4-41)
H idc idc , x s
Gidc + 1
z
= (4-42)
Gidc 2nDdc , x
The right-hand plane zero frequency ωRHP introduces an additional 90° phase shift,
while adding a +20dB/decade slope. However, the closed-loop current controller allows
for the transfer function to have a dominant pole that coincides with the current transfer
For most converters ωRHP>>ωz, and the bandwidth will be determined by the dominant
pole at ωz and the 0dB crossover frequency for the voltage controller, ωv,0dB can be found
from (4-27) and (4-42), as long as the gain is more than 0dB:
Gvc (4-44)
v,0 dB = z
Gidc
The proportional gain of the voltage controller can be calculated using the desired
v ,0 dB ,CL (4-45)
k p ,vc =
v ,0 dB
Since ddc,x is calculated by the current compensator, the converter delay is already
taken into account. Moreover, the crossover frequency of the voltage control loop is much
98
Modulation and control system of BMCIs
lower than the sampling frequency and the compensator zero can be set an octave below
ωz. This is a sufficient margin that avoids excessive overshoot at lighter loads:
z (4-46)
ki ,vc = k p,vc
2
The Nyquist plot of the forward transfer function must be evaluated, as the right-hand
plane zero can cause an instability if its frequency ωRHP is too close to ωv,0dB,CL. The control
system can be evaluated only at peak loading, as according to (4-43) ωRHP increases with
frequency. The open loop crossover frequency ωv,0dB will remain relatively constant, as
according to (4-27) the capacitor transfer function dominant pole frequency is inversely
as long as RSM>>rarm, where rarm=rL+nrSM. The open loop plant and compensator transfer
function is:
99
Chapter 4
Phase ( o)
0o
Frequency (rad / s)
o
90
180o
270o
Figure 4-12 Simplified bode plot of voltage closed loop, open loop, and compensator transfer functions
voltage differences between the top and the bottom converter arms. This can be avoided
by introducing an extra compensator, that maintains equal mean values of the top and
bottom arm of the same leg. The error signal, that needs to be maintained at 0 is:
injected. To avoid it flowing through the 3-phase load, this current needs to be common
to both converter arms in the same leg. This can be achieved by adding another circulating
current term, ibal,x. The power transfer utilises the differential duty cycle dac,x, and ibal,x
100
Modulation and control system of BMCIs
In equation (4-50) the ac duty cycle has been normalized to the amplitude of the ac
duty cycle, which allows for constant controller gain over different values of the
modulation index m. The balancing function can be modelled by the following transfer
function:
dac, x,rms
vC ,a /b, x ( s ) = ibal , x ( s ) (4-51)
sC
The balancing transfer function only has a pole at the origin, which means it can use a
controller, the arm balancing control system should be unconditionally stable, as long as
its crossover frequency is a decade below the circulating current loop bandwidth.
Since this system can use a proportional controller, its output must be multiplied by
the ac duty cycle of each phase and this is achieved by multiplying its output by a
The controllers discussed in section 4.2.2 are shown in a single block diagram in Figure
4-13:
101
Chapter 4
vC,a,x 3 + 3
-
3
3
vC,b,x
ibal,x*
iref,x*
+ ddc,x
+ 3 +
vC*
- -
3 3
vC,x idc,x
To implement 3-phase motor control the BMCI can use well established controllers, as
the ac side of the converter is independent of the dc side, as long as the peak sum ddc,x +
dac,x < 1. For a traction drive, the control algorithm must implement field weakening to
fully utilise the maximum power available from the converter and the prime mover, as
well as third harmonic injection to achieve the maximum line-to-line voltage the
A key property of the asynchronous ac motor is the difference between rotor rotating
frequency ωr and 3-phase supply frequency ωe, with a slip frequency ωslip when the
p (4-52)
e = slip + r
2
When the motor is fed by a balanced 3-phase power supply, each phase will have the
same current magnitude displaced by 2π/3. At the steady state, the machine phase
102
Modulation and control system of BMCIs
rs Llk,s Llk,r
+ Rr
vac,x iac,x,s LM i
ac,x,r s
, where ix,s is the stator current and ix,r is the reflected rotor current. Each machine phase
has a stator inductance Llk,s, stator winding resistance rs, magnetizing inductance LM,
primary reflected leakage rotor inductance Llk,r and resistance Rr, which is dependent on
e − r (4-53)
s=
e
The big advantage of vector control is that it can control an ac machine with similar
torque dynamics as a dc-machine controller, while operating with a fixed PWM frequency.
This can be done by expressing the asynchronous machine currents as a vector that is
The Clark and Park transformations are standard tools used to convert the machine’s
the synchronous frequency ωe. For a matrix of 3-phase voltages vx = [vu, vv, vw], they can be
103
Chapter 4
1 1
1 − − vu
2 2 2 (4-54)
v = vv
3 3 3
0 2 − 2 vw
1 1
1 − 2 − 2
Clarke = (4-55)
3 3
0 2 − 2
transformation can be simplified for calculating iac, α and iac,β. The Clark transformation is
iac , = iu
(4-56)
2
iac , = ( iv − iw )
3
Since the terms in the equations are also oscillating at ωe, an additional transformation
is required to refer the quantities to the rotating frame of reference. The park transform
is used to calculate the direct and quadrature 2-phase rotating voltage matrix
Using the conversion matrix in (4-58) the d and q currents can be calculated using
idq = ΓParkiαβ:
104
Modulation and control system of BMCIs
The per-phase model of the asynchronous motor can be derived from the per-phase
Applying the matrices ΓClarke and ΓPark to (4-60), the per-phase stationary frame circuit
d L d (4-61)
vac ,s ,d = rs iac ,s ,d + Llk ,s iac ,s ,d + M r ,d − e Llk ,siac ,s ,q
dt Lr dt
d L (4-62)
vac ,s ,q = rs iac,s ,q + Llk ,s iac ,s ,q + e M r ,d + e Llk ,s iac ,s ,d
dt Lr
, where Φr,d is the rotor d axis flux:
Lr = LM + Llk ,r (4-64)
p L (4-65)
Tem = r ,d M iac,s ,q
2 Lr
105
Chapter 4
d L i (4-66)
r ,d = M ac , s ,d − r ,d
dt r r
The slip frequency can be calculated from the rotor flux and stator current iac,s,q:
LM iac , s ,q (4-67)
slip =
r ,d
The electrical angle, used for the Clarke and Park transforms uses the rotor speed ωr
The stator current models of the d and q axis, have cross-coupling components,
containing the rotor flux and rotor flux derivative. The derivative of Φr,d is considered to
be negligible and the control system can be compensated by the simplified terms:
LM d
vcomp ,q = e r ,d + e Llk ,s iac ,s ,d (4-70)
Lr dt
Adding (4-69) and (4-70) to (4-61) and (4-62) respectively, results in the following
d L d (4-71)
Llk ,s iac ,s ,d = −rs iac ,s ,d − M r ,d + vcomp ,d
dt Lr dt
d L d
Llk ,s iac ,s ,q = −rsiac ,s ,q − e M r ,d + vcomp ,q (4-72)
dt Lr dt
Since the machine voltages are generated by the BMCI, the d and q stator voltages can
be replaced by:
106
Modulation and control system of BMCIs
vac ,s ,d = d ac ,d nVC
(4-73)
vac ,s ,q = d ac ,q nVC
The d and q duty cycles, dac,d and dac,q, are generated by the outputs of the d and q axis
To calculate the stator current compensator gains, kp,iac and ki,iac, the transfer function
of the motor is first required. As the BMCI is used in a traction converter, the rotor flux
dynamics will be determined by the asynchronous motor’s inertial load. For a vehicle
application, the dynamics of the mechanical system will be several orders of magnitude
slower than those of the electrical system. Thus, the flux dynamics are ignored for the
small-signal model and it can be derived by substituting (4-73) in (4-71) and (4-72), and
rarm
rs + (4-76)
z ,s = 2
Llk , s
nVC
Giac = (4-77)
r
rs + arm
2
The crossover frequency of the stator current plant can be calculated from the dc gain
107
Chapter 4
, where βiac is the ac current feedback gain. As with the dc current controller, the closed
loop bandwidth ωiac,0dB,CL can be set by shifting the plant’s crossover frequency using the
The integrator gain can be set relative to ωiac,0dB,CL to give a phase margin of 74°:
The motor plant current transfer function is of the first order and the maximum phase
shift is 90°. As with the dc current controller, the modulator delay is taken into account
The final step of the vector controller is to calculate the three ac duty cycles dac,x that
are used in the generation of the arm inserted voltage commands. This is done by
executing the inverse of the transformations described in section i. The inverse Park
108
Modulation and control system of BMCIs
d ac , s ,u = d ac , s ,
−d ac , s , + 3d ac , s , (4-83)
d ac , s ,v =
2
−d ac , s , − 3d ac , s ,
d ac , s , w =
2
4.2.3.2. Third harmonic injection and control system overview
inverter’s dc-link. Triangular third harmonic injection uses the 3-phase ac duty cycles,
dac,s,x, to calculate a common mode signal, that does not require the phase angle θe.
The resultant waveform is a triangular, and is oscillating at 3 times the frequency ωe,
and the peak phase-to-phase voltage is increased by 15.4% to achieve full utilisation of
the available voltage from the capacitor sub-modules. The peak value occurs at values of
In Figure 4-15, the modulation duty cycles of each phase are shown and they are
calculated by subtracting dthi to the three values dac,s,x to calculate the duty cycles dac,x:
109
Chapter 4
The block diagram is shown in Figure 4-16. The Clarke and Park transformations have
been lumped into a single abc/dq block, while the inverse Park and Clarke
ωr
-
ωr* + iq*
idq*
idq - + - + dac,x
abc + dq
+
iuvw 3
dq abc max 0.5
-
+
θe ωe min
Kiac ωe
iq 1 ωslip 1 θe
r s
id
ωr pp
τr
Φr,d
LM
According to equation (4-66), the flux can be modelled as a first order system with the
input being the current id. This allows for the flux to be controlled by using the estimated
flux as the feedback variable and id as the input. The flux controller bandwidth must be at
110
Modulation and control system of BMCIs
When the BMCI is working with modulation index < 1, the available duty cycle, equal
to 1-max(dac,x)-ddc,x, can be used to balance the voltage between each pair of arms a and b.
This balancing duty cycle dlf is differential to arms a and b, the same as dac,x. For the
balancing power to not flow through the load, dlf is common to all three legs:
d a / b ,u = d dc ,u d ac ,u dlf
(4-86)
d a / b ,v = d dc ,v d ac ,v dlf
d a / b , w = d dc , w d ac , w dlf
d b ,u − d a ,u d b ,v − d a ,v
vuv = − = d ac ,u nVC
2 2
d b ,u − d a , v d b , v − d a , w (4-87)
vvw = − = d ac ,v nVC
2 2
d b , w − d a , w d b ,u − d a ,u
vwu = − = d ac ,w nVC
2 2
Power transfer between arms a and b is achieved by injecting an additional circulating
current ilf,x, common to both arms, that is in phase with the common mode voltage, thus
discharging the capacitors of one arm and charging those of the other one. The current ilf,x
is specific to each phase leg, and unless all three currents sum up to 0, the compensation
The power plf,x must be alternating to achieve stability with capacitors in the sub-
lower than the transistor switching frequency fQ, but high enough to cause a small ripple
controlled by the arms’ switching functions, its rate of change is only limited by the
switching speed of the transistors in each sub-module. This allows for the shape of dlf to
111
Chapter 4
In an ideal case, the balancing current ilf,x would be also have a rectangular waveform.
ilf,x*
vC,x,p 3 + 3
3
-
3
vC,x,n
The compensator in Figure 4-17 is a proportional type, as a dc value of ilf will result in
a voltage ramp across the capacitors. The gains can be set according to the model of the
dlf
vC ,a / b, x ( s ) = ilf , x ( s ) (4-90)
sC
The low-frequency balancing controller has similar plant model to the normal range
arm balancing controller, and thus can use the same structure, with just a change in the
The control systems for the full converter are grouped in two parts – dc-side and ac-
side. The dc-side control system includes the circulating current controller, balancing
ddc,x 3 + 3 dx,a n n
dx,a,i
-
n n
vC,x,a ka,x,i
min
+ + dx,b
dac,x dd c,x + + n n
-dd c,x
dx,b,i
3
n n
vC,x,b kb,x,i
The ac and dc duty cycles ddc,x and dac,x must be complementary, but they control
separate variables. To avoid interference between the dc-side and ac-side current
For a vehicle, there will be two energy sources – the prime mover, connected to the dc-
side of the converter, and the kinetic energy stored in the vehicle’s mass. If the duty cycles
When the ac duty cycle causes the arm duty cycles to saturate, for a BMCI with positive
dc-side voltage, this will always occur at da/b,x=1. On the negative half-cycle, the converter
will not be in saturation until dac,x≥1+ddc,x and at dac,x=ddc,x-dac,x,max the total inserted dc
voltage will no longer be 2ddc,xnVC as shown in Figure 4-19. This offset can cause a very
113
Chapter 4
Figure 4-19 Effect of arm duty cycle saturation (above) on filtered arm voltage (below)
If the dc power source cannot absorb power under any circumstances, an unlimited
closed-loop circulating current controller will keep increasing ddc,x until a value of 1 is
reached. This will raise the value around which dac,x is oscillating, clipping it at dac,x=1-
ddc,x,max. This asymmetry will occur during the next half-cycle as well, just in the other
converter arm. The end result is a clipped ac voltage waveform. In a large traction motor,
the winding resistance can be very small, and even a change of a few volts can cause a
large current to flow, which in turn will apply a very high torque to the traction motor,
Considering the two scenarios the strategy adopted is to use a variable limit for the ac
An important point is that the limit will clip dac,x symmetrically. This strategy ensures
that each leg can return power to the dc source or dissipate it in a braking chopper
114
Modulation and control system of BMCIs
resistor. The limits use the maximum value of the dc duty cycles of the whole converter to
4.3. SUMMARY
This chapter reviews the different modulation schemes used for MCCs and are
applicable to the BMCI. The control systems required for the operation of the converter
are also presented, together with the design methodology used for tuning.
Pulse-width modulation is the type used for this thesis, as it is readily available from
equal switching losses across all sub-modules, without intermittent peaks and troughs
solution does not introduce an error between the arm command and the actual average
The BMCI current and voltage controllers are identical to those of a cascaded boost dc-
dc converter and 3-phase inverter, and the different break frequencies required for the
115
Chapter 4
116
General design for an n-module BMCI and experimental converter
This chapter describes the design procedure for an n-module boost multilevel
cascaded inverter (BMCI). The relationships between number of sub-modules and ac-side
voltage quality, and size of leg inductance are examined, as well as the effects of different
modulation strategies. The converter design uses dc-side voltage, ac-side peak voltage,
machine base frequency, and maximum capacitor voltage ripple as design parameters and
the presented method is used to design the laboratory prototype, which is used in this
A design example for a fuel-cell BMCI drive can be found in Appendix A. The schematics
arm n, the total leg inductance LLeg, and the submodule capacitance CSM. The number of
sub-modules has a direct effect on the other two parameters and must be defined first.
The capacitance and the inductance are then determined by the maximum allowable
Following the same conventions as Chapter 3, the general converter diagram is shown
in Figure 5-1:
117
Chapter 5
SMa,x,1 ia,x
Ca,x,1
SMa,x,n va,x x = u, v, w
Ca,x,n
L
w
Vdc vLL v N
x=u ix=u
SMa,x,1 L
Cb,x,1
vb,x
SMa,x,n
Cb,x,n
ib,x
The converter’s dc-side voltage is vdc, top arm inserted voltage is va,x, bottom arm
inserted voltage is vb,x, where x is the phase designator, equal to u, v, or w. The sum of each
pair of converter arms produces a dc voltage vdc,x, and the difference produces the phase-
vb, x − va , x (5-2)
vx =
2
For an ideal converter the inserted dc-side voltage vdc,x → vdc, and the phase voltage vx
is a sinusoid with amplitude Vx, and for a 3-phase output each leg will have a displacement
vx = Vx sin( x t − x ) (5-3)
118
General design for an n-module BMCI and experimental converter
va , x = vdc, x − v x
(5-4)
vb, x = vdc, x + v x
The top and bottom arm currents are ia,x and ib,x. The sum of these two currents results
in the leg circulating current idc,x, the difference forms the line current, and the sum of the
iIN = i
x =u , v , w
dc, x
(5-7)
The modular cascaded converter (MCC) was developed for use in high-voltage dc
(HVDC) substations. With transmission line voltages in the range of tens to hundredths of
kVs, there is no single device that can withstand the full dc-link voltage, and a high number
In comparison traction drives in the railway sector work with peak ac-side voltages
between 750V and 3.5kV, requiring IGBT collector-emitter voltages rated between 1.2kV
and 6.5kV. In the automotive market the choice of transistors increases even further when
Thus, for a low-voltage MCC-type converter, the required ac-side voltage is not the
only design parameter, that determines the number of sub-modules n, and factors such as
leg inductance and ac-side voltage distortion become more dominant. High values of n
119
Chapter 5
however increase the converter’s complexity, as more gate drivers and voltage
The maximum dc-side voltage and the peak line-to-line output determine the
, where VLL is the amplitude of the line-to-line voltage, and VIN is the maximum dc-side
voltage. This value is for a converter using third harmonic injection, as sinusoidal PWM
does not utilise the full voltage available from the capacitors. Equation (5-8) shows that as
the converter dc-side voltage reduces, the required maximum arm voltage relaxes. The
max( va / b, x ) (5-9)
n= ,
VC ,nom
, where VC,nom is the nominal sub-module capacitor voltage. The arm in the (BMCI) uses
full-bridge (FB) sub-modules, and can produce both positive and negative voltages. The
number of voltage steps in each arm, narm, is 2n+1 as a maximum, which is achieved at dc
duty cycles ddc<0.5/n. With every increment of ddc by 0.5/n, narm decreases by 1. The
number of phase voltage levels, nx, is dependent on narm, as vx is derived from (5-2). It
should be noted that the voltage vx is measured between the load neutral and the common
(HB) sub-modules, the number of output levels for each arm, narm is n+1, with ddc=0.5. For
120
General design for an n-module BMCI and experimental converter
, where round() is a rounding function. The output of (5-10) is independent on the type of
Chapter 4 examines the control implications of two modulation schemes - (PSC) and
n n
n da,x n da,x
2 2
n n
1 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 2
n n
n n
n n
θcarrier θcarrier
n n
n n
2 2
n n
1 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 db,bx 2 db,bx
n n
n n
n n
a.) b.)
Figure 5-2 Modulation for top and bottom converter arms, Phase shifted carrier a.), and Phase disposition carrier b.)
Each carrier in Figure 5-2 a.) oscillates at frequency 2fsw,Q, which is twice the individual
transistor switching frequency fsw,Q, while the carriers in Figure 5-2 b.) oscillate at nfsw.
However, the effective arm switching frequency is the same for both cases, as the PSC
carriers are interleaved, and produce the same switching frequency farm=nfsw.
To compare the effects of the two modulation schemes more easily, the PSC
121
Chapter 5
produce the same arm voltage waveform [151]. APOD is simpler to analyse, as each
modulation carrier corresponds to a single level and oscillates at the effective arm
TSW,eff TSW,eff
n
da,x n da,x
dlocal,a,x 2 dlocal,a,x
n
1
ddc,x n ddc,x
0
Time Time
1
n
2
n
n
n
θcarrier θcarrier
n
n
2
n
1
ddc,x n ddc,x
0
Time Time
dlocal,b,x 1 dlocal,b,x
n
db,bx db,bx
2
n
n
n
a.) b.)
Figure 5-3 PWM modulation using APOD carriers a.) and PDC carriers b.)
With each carrier localised at a certain arm level, the PWM waveform will have a local
duty cycle dlocal,a/b,x, which cycles from 0 to 100% every time the arm modulation signal
increases or decreases by 1/n. When the arm duty cycle becomes negative, the local duty
cycle is the fraction of the switching cycle. The voltage is most negative during the active
Since each arm has its own set of carriers, arm a and arm b carriers can be offset by an
angle ϑcarrier, also shown in Figure 5-2 and Figure 5-3. The effect of ϑcarrier is more
pronounced at values of ddc,x that are divisible by 0.5/n, where the local arm duty cycles
122
General design for an n-module BMCI and experimental converter
are complementary. When a rising edge from arm b aligns with a falling edge from arm a,
the voltage step of vx is equal to VC,nom and has a switching frequency fsw,eff. In the opposite
condition, the two pulses are effectively interleaved and the resultant vx has voltage steps
of VC,nom/2, with switching frequency 2fsw,eff. An example, using PSC modulation and ddc,x =
n n
n da,x n da,x
2 2
n n
1 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 2
n n
n n
n n
n n
n n
2 2
n n
1 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 2
n dac,x n dac,x
n vx/Vc n vx/Vc
n n
a.) b.)
Figure 5-4 Pulsed waveforms of arm and phase voltages, with carrier phase shift of 180° a.) and 0° b.)
123
Chapter 5
The relationship between ac-side voltage levels and phase shift is also different for odd
or even number of sub-modules. This is due to the position of ddc,x relative to the local
carrier. The relationship between n, ϑcarrier, and ddc,x is summarised in Table 5-1:
Table 5-1 BMCI phase voltage levels with different carrier phase shift
carrier = 0 carrier =
2n
Number of sub-
n
modules
Number of phase
Number of phase
…, n
Since the dc duty cycle ddc introduces an offset, around which the sinusoidal dac
oscillates, over the range from 0 to 1 there are 2n operating points, where the modulated
signal produces symmetrical waveforms in each arm. These are the boundary conditions
described in Table 5-1. For values of ddc between these boundaries, the arm pulses are not
symmetric between the top and the bottom arms, but this asymmetry repeats each half-
cycle, making the ac-side voltage free of second order harmonics. At the phase output, this
gives rise to additional switching actions, resulting in a waveform that oscillates between
3 levels over one carrier cycle. Compared to a conventional multilevel inverter, i.e. floating
124
General design for an n-module BMCI and experimental converter
capacitor or diode-clamped inverter, the phase voltage always switches between two
neighbouring levels:
n n
n da,x n da,x
2 2
n n
1 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 2
n n
n n
n n
n n
n n
2 2
n n
1 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 2
n dac,x n dac,x
n vx/Vc n vx/Vc
n n
a.) b.)
Figure 5-5 Arm and phase voltage waveforms with 0.5/n < ddc,x < 1/n, and ϑcarrier = 180° a.) or ϑcarrier = 0° b.)
The two normalised phase voltage waveforms, from Figure 5-5, are both formed of half
and full-steps. A judgment cannot be made as to the inverter levels. Since the dc duty cycle
is between two boundary conditions, the performance is expected to be “in between” the
effective output levels, and switching frequency. To better compare the inverter harmonic
125
Chapter 5
performance the Total Harmonic Distortion (THD) is evaluated for several cases, using
100
1 (5-11)
THD =
V1
V
i = 2 , 3,...
i
2
The THD is evaluated for a converter with n=3, in 4 cases, by keeping the modulation
index at maximum, M = 1-ddc,x. The dc duty cycle ddc,x is increased by 0.1/n from 1/n to
1.5/n:
35
30
25
20
% THD
PSC_0
15 PDC_0
PSC_PI
10
PDC_PI
5
0
0.3 0.35 0.4 0.45 0.5
Ddc
Figure 5-6 THD results at maximum modulation index, for PSC, ϑcarrier= 0°, PDC, ϑcarrier= 0°, PSC, ϑcarrier= 180°, PDC,
ϑcarrier= 180°
At the two boundary conditions the THD is equal for PSC and PDC modulation, when
ϑcarrier takes opposite values. As ddc,x increases, the number of arm voltage levels goes
down, causing harmonic distortion to increase. However, when ϑcarrier = 0 for the PSC
Overall, the BMCI can achieve a higher number of levels, than the buck MCC, for the
inverter with the same number of levels. For example, a converter with n=2, operating
with a voltage boost of 2, i.e. ddc=0.33, can have between 4 and 7 voltage steps. The highest
126
General design for an n-module BMCI and experimental converter
number of voltage steps is achieved when ddc→0, but this results in voltage gain GV→∞
converters are designed for voltage gains between 2 and 5, as is the experimental
The value of ddc,x can be chosen based on the required voltage gain:
2 2VLL (5-12)
d dc , x =
VIN
The inductor design for 2-level boost converters is a function of boost ratio, dc-side
voltage, system power, and switching frequency. The process is simple, as under any
steady-state condition, the inductor voltage has constant amplitude and frequency.
In a BMCI the inductor voltage varies over the cycle of the ac-side voltage waveform,
in terms of both frequency and amplitude. For this reason, analysis is necessary to find
the highest volt-seconds product, that will give the largest current ripple.
The converter’s 6 inductances limit the arm current ripple and are necessary to
achieve boost operations. As they carry dc current, the inductors need to have an air gap,
or be made of low permeability material. In any case, the size and the weight of the
inductor are proportional to the energy stored. The choice of inductance value is made by
evaluating the maximum volt-seconds across the leg inductor, and the peak current ripple,
Δidc,x:
127
Chapter 5
1 TSWeff v + v
idc, x = max vIN − a , x b, x dt (5-13)
Lleg 0 2
As the dc-side voltage vdc is a constant, the inductor volt-seconds are a function of the
sum of the arm voltages. This is similar to the instantaneous phase voltage, and the phase
Using the same arm voltage waveforms from Figure 5-4, the voltage vdc,x for ddc,x=1/n
128
General design for an n-module BMCI and experimental converter
n n
n da,x n da,x
2 2
n n
1 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 2
n n
n n
n n
n n
n n
2 2
n n (va + vb) /Vc
(va + vb) /Vc
1 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 2
n n
n n
n n
a.) b.)
Figure 5-7 Instantaneous arm voltages and resultant sum for ϑcarrier = 180° a.) and ϑcarrier = 0° b.)
When ϑcarrier = 0° the two waveforms are complementary and sum-up to a constant
value, resulting in cancelation of the switching waveforms. With ϑcarrier = 180° the summed
dc voltage vdc,x = va,x + vb,x is switching by voltage steps of ±VC around the nominal point.
These oscillations appear for both cases of ϑcarrier when ddc,x is not divisible by 1/n. The
difference between the carrier angles is not as well defined for an example with 1/n < ddc,x
< 2/n:
129
Chapter 5
n n
n da,x n da,x
2 2
n n
1 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 2
n n
n n
n n
n n
n n
(va + vb) /Vc (va + vb) /Vc
2 2
n n
1 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n Edge-alignment
instance
2 2
n n
n n
n n
a.) b.)
Figure 5-8 Arm and dc voltage waveforms with 0.5/n < ddc,x < 1/n, and ϑcarrier = 180° a.) or ϑcarrier = 0° b.)
The dc voltage now oscillates either between two levels, VC/n and 2VC/n, or between 3
levels – 0, VC/n, and 2VC/n. In Figure 5-8 b.) the point at which the top and bottom arm
switching edges align has been highlighted, where the arms a and b’s waveforms switch
with the same polarity. The calculation of the exact volt-seconds product is difficult by
130
General design for an n-module BMCI and experimental converter
In the examples so far, the ratio of effective switching frequency to modulated signal
frequency, kf = 2πfsw/ω, has been kept low, so that the pulses are clearly visible. Under
this condition, however, the local duty cycles do not pass through all possible values
between 0 and 100%, due to the discrete nature of the modulation. The effect of utilising
Δt
n n
n n
2 Δ dac 2
n n
1 2 dac 1
n ddc,x n ddc,x
0 0
Time Time
1 1
n n
2 2
n n
n n
n n
dx,a/b dx,a/b
n n
n n
2 2
n n
1 1
n π n π
n n
0 0
Carrier Carrier
1 angle 1 angle
n n
2 2
n n
n n
n n
Figure 5-9 Construction of arm duty cycle trajectory for APOD (left) and PCD(right) modulation
By establishing that kF→∞, the intersection between the modulation signals of arms a
and b, and the respective local carrier, occur at time intervals Δt→0. By increasing the ac
control signal dac,x and overlaying the point of each leading edge in the effective switching
period, the actual trajectories of the two arm duty cycles can be traced. When the arm a
and b duty cycles align vertically, i.e. the switching edges occur at the same carrier angle,
131
Chapter 5
either a summation or subtraction occurs. The direction of the switching edge is inferred
from the slope of the local duty cycle: positive cycle (increasing with carrier angle) results
In Figure 5-10 the effects of ddc,x and ϑcarrier are shown in 4 particular cases: complete
pulse cancellation in vdc,x for ϑcarrier=0, complete pulse cancellation for ϑcarrier=180°, edge
coincidence for ϑcarrier=0° and ϑcarrier=180°, with 0.5/n < ddc,x < 1/n:
Figure 5-10 Duty cycle trajectories for PSC and APOD modulation
With ddc,x is not a multiple of 1/n or 0.5/n, the pulse edges of the two converter arms
only align once per increment of dac,x by 1/n. When the direction of the edges coincides,
the resultant vdc,x pulse has twice the amplitude, i.e. 2VC, and a switching period of
Tsw,eff=1/fsw,eff. At other values of dac,x the waveform of vdc,x switches either between 2 or 3
levels, but will have smaller volt-seconds product over a period of Tsw,eff. Using Figure
132
General design for an n-module BMCI and experimental converter
d coinc = nd dc, x − floor (nd dc, x ), carrier = 0
0.5 (5-14)
d coinc = nd dc, x − round (nd dc, x ) + 0.5, carrier = 180 , d dc, x
n
d dc, x + 0.5 0.5
d coinc = , carrier = 180 , d dc, x
n n
As the arm local duty cycle repeats over every change by 1/n, the instantaneous
inductor voltage duty cycle becomes only a function of the ac component of vdc,x, with
peak-to-peak value of either VC or 2VC. The ripple current amplitude, at point of edge
coincidence, is:
The key difference in the PDC case is that complete pulse cancellation only occurs
when ϑcarrier=180° as only that angle allows for symmetric switching actions for the two
133
Chapter 5
arms. For ϑcarrier=0 there are pulse coincidences, resulting in pulses with amplitude of 2VC
twice per level. The peak current ripple will be caused by the higher of the two
coincidence duty cycles. When ϑcarrier= 180°, and ddc,x is not divisive by 0.5/n, there are no
coincidences between the top and bottom arm pulses and the peak volt-seconds product
will occur when one arm voltage has no pulsed component, while the other one still does.
At these points the voltage step is equal to VC, and the switching period is Tsw,eff. The duty
cycles that causes the highest current ripple for the PDC case are:
2VCTsw,eff max ( dcoinc1 (1 − dcoinc1 ) , d coinc 2 (1 − d coinc 2 ) ) , carrier = 0
1
idc , x =
(5-17)
Lleg
1
idc , x = dcoinc (1 − d coinc )VCTsw,eff , carrier = 180
Lleg
The design method proposed in this work chooses the inductance value based solely
on the peak current ripple amplitude, over one modulation waveform cycle. Another
important aspect is the frequency spectrum of the converter current, as this has direct
effect on the electro-magnetic interference (EMI) filters required for a finished product.
The frequency spectrum of the circulating current, of one phase leg, is examined
Value:
Parameter
LLeg 100uH
fsw 1kHz
VC 300V
n 3
ddc,x 0.45
The dc current level is removed to accentuate the spectrum of the ac components. The
Figure 5-12 Circulating current spectrum for PSC and PDC modulation, ϑcarrier = 0°
135
Chapter 5
The spectrum of the PDC case has a very strong component at the arm effective
switching frequency fsw,eff. For PSC modulation, the harmonics with the highest amplitude
are also at same frequency, but with lower value. The lower ripple for the PDC case is also
reflected in the calculating peak current ripple, using (5-15) and (5-17). The results are
355A for the PSC case, and 370A for the PDC one. This ripple amplitude values are very
similar, and even though the peak amplitude of the PSC case is higher, the PDC scheme’s
By changing ϑcarrier to 180° it is expected that the current ripple will reduce, as ddc,x is
Figure 5-13 Circulating current spectrum for PSC and PDC modulation, ϑcarrier = 180°
The peak current ripple calculated is 355A for the PDC case and 258A for the PSC
scheme. Like the previous case, the scheme with the lower harmonic amplitude has the
higher peak current ripple, but the second harmonic, of the arm effective frequency, is
identical.
136
General design for an n-module BMCI and experimental converter
Even though the peak current ripple for PDC modulation with ϑcarrier = 180°, is as high
as the ϑcarrier = 0° case, the harmonic spectrum is less populated, and would consequently
be easier to filter.
The physical design of the leg inductors is based on the total energy stored in them.
The inductor energy is dependent on the peak arm current, which is:
L(max (ia / b, x ))
2
(5-18)
WL ,leg =
2
Individual leg inductors increase the BMCI’s output inductance. However, a traction
motor has a relatively high leakage inductance, and any additional phase inductance is
undesirable.
If each pair of leg inductors are lumped into a single centre-tapped inductor, with high
coupling factor between the two windings, both inductor energy and output inductance
can be lowered, as the magnetic flux established in the core will be only due to the dc
component of the current. The ac components create effectively cancel out. The size of the
coupled inductor only depends on the peak circulating current idc,x, and is not affected by
the ac output current iac,x. The two possible inductor configurations are shown below:
137
Chapter 5
ϕa,x
Lx
La,x
La,x
Lx
{ La,x,lk
LM,x
ϕM,x
ϕb,x
Lb,x
Lb,x
{ LM,x
Lb,x,lk
Figure 5-14 Circuit and core flux for two separate leg inductors (left) and a single centre-tapped inductor (right)
In the case of the centre-tapped inductor, the two windings will not have a unity
coupling factor and there will be some leakage flux for each one. This is accounted for by
the leakage inductance, La/b,x,lk, which however does not influence energy stored in the
core air gap. It does appear, however at the output of the converter as Lout,x and is added
For the two separate leg inductors case, the values are calculated by using:
Lleg (5-19)
La , x = Lb, x =
2
2
L I (5-20)
WL , x ,max = 2 a / b, x I dc, x + ac, x
2 2
La / b, x (5-21)
Lout, x =
2
For the single centre-tapped inductor, the value of the magnetizing inductance LM, i.e.
138
General design for an n-module BMCI and experimental converter
Lleg (5-22)
LM , x
4
4 LM , x (5-23)
WL, x ,max =
2
I dc, x
2
La / b, x ,lk (5-24)
Lout, x =
2
The magnetizing inductance LM,x should be slightly smaller than the value calculated
in (5-22), as the leakage inductances are added to the total leg inductance.
Energy stored in each arm depends on vdc,x, vx, fundamental load frequency ω , voltage
boost ratio GV,dc, and instantaneous arm powers px,a and px,b. The converter voltage gain is
controlled by the dc duty cycle ddc,x. The steady-state output conditions of the BMCI are:
(5-25)
vx (t ) = Vx sin(t ) = m (1 − ddc , x ) n vC sin(t )
(5-26)
vdc, x (t ) = Vdc = 2d dc, x n VC
, where m is the modulation index, φ is the current displacement angle, and Ix is the
magnitude of the output current. The design parameters are the total arm capacitance
Carm, the capacitor voltage VCx,a/b, and the total arm energy, which are given by the
following equations:
139
Chapter 5
Carm =
C (5-28)
n
(5-29)
VCx,a / b = nVC ,nom
wx ,a / b =
1 2
Carmv(t ) Cx (5-30)
,a / b ,i
2
Since the output power is pulsating at the load frequency, the stored arm energy varies
over one cycle, with peak fluctuation, Δwc, causing a voltage ripple of ΔVCx,a/b on top of the
average dc voltage:
2 (5-31)
= 4 Carm VCx ,a / b vCx ,a / b
From (5-31) the minimum capacitance, for achieving a certain ripple, becomes:
wx , a / b
Carm (5-32)
4vCx, a / b VCx, a / b
The peak energy fluctuation can be found by integrating the instantaneous arm power:
T T
w x ,a / b = max p x ,a / b (t )dt − min p x ,a / b (t )dt
(5-33)
0 0
, where the instantaneous arm power can be obtained from the arm voltages and currents:
140
General design for an n-module BMCI and experimental converter
V I
px ,a / b (t ) = dc , x Vx sin (t ) I dc , x x sin (t − )
2 2
V I V I sin(t − )
= dc , x dc , x dc , x x Vx I dc , x sin(t )
2 4 (5-36)
0.5 Pdc , x
Vx I x VI
− cos( ) + x x cos(2t − )
4 4
0.5 Pac , x
According to (5-36), the top and bottom arm powers differ in the oscillating
components, that are in counter phase. Thus, only the top arm power is used for the
following the analysis. The arm power contains two dc-components and 2 oscillatory
components - at the fundamental and twice the fundamental frequency. The only
difference between arms a and b is the phase of the fundamental component and further
analysis will only examine the power of one converter arm. Over time the net capacitor
power must be 0, and by using Pac,x = Pac,x, the arm power can be reduced to:
Vdc, x I x sin(t − ) VI
px,a (t ) = Vx I dc, x sin(t ) + x x cos(2t − ) (5-37)
4 4
The dc and phase voltage amplitudes in each arm are produced by the converter’s
capacitor voltages. Substituting with (5-25) - (5-27) in (5-37), the arm power can be made
a function of the average capacitor voltage, the dc duty cycle ddc,x, and the modulation
index m:
141
Chapter 5
(1 − d )
2
d dc , x n VC I x
px ,a (t ) = sin(t − ) − m n VC I x cos( )sin(t ) +
2 dc , x
(5-38)
2 2dcirc , x
m (1 − d circ , x ) n vC I x
+ cos(2t − )
4
Additionally, by defining the apparent power SOUT and the voltage boost ratio as:
m (1 − d dc , x ) n vC I x
SOUT , x = Vx ,rms I x ,rms =
2 2
(5-39)
1 − d circ , x
G
V , dc =
d circ , x
SOUT , x
px ,a (t ) = sin(t − ) − m SOUT , xGV ,dc cos( )sin(t )
MGV ,dc
(5-40)
SOUT , x
+ cos(2t − )
2
There are two power components in(5-40), that can be denoted as S1,x, with phase ϕ1,x,
( )
2
S cos( ) 1 − mG 2
SOUT , x
2
S1, x = + sin( )
OUT , x V , dc
mG V , dc mG
V ,dc
(5-41)
sin( ) tan ( )
1, x = tan −1 = tan −1
cos( ) (1 − mGV2 ,dc )
1 − mGV2 ,dc
SOUT , x
S2, x = (5-42)
2
2, x =
Substituting (5-40) - (5-42) into (5-33), the peak energy fluctuation for arm a becomes:
142
General design for an n-module BMCI and experimental converter
S1, x
cos ( t − 1, x ) + sin ( 2t − 2, x )
S2, x
wx , a = max −
2 (5-43)
S1, x
cos ( t − 1, x ) + sin ( 2t − 2, x )
S2, x
− min −
2
Evaluating the peak arm energy, for a given frequency can now give an arm
capacitance value by substituting (5-43) in (5-32). For a traction converter, the submodule
capacitance should be designed according to the peak energy fluctuation over the full
speed range of the converter, including field weakening, where the converter is operating
at constant output power. This can be achieved by ramping output frequency, output
power, and dc-gain up to base speed, and then keeping power and voltage gain constant,
while increasing frequency, simulating the constant torque and power regions:
As an example, the peak energy fluctuation was evaluated for a laboratory prototype,
The converter arm energy fluctuation is simulated and a surface plot of the resulting
peak-to-peak ripple, as a function of output frequency and power factor, can be seen in
Figure 5-15.
143
Chapter 5
Figure 5-15 Sub-module arm energy fluctuation for experimental prototype converter, with value at the converter’s
operating point
current. However, the ripple-reduction control system can provide full compensation at
low frequency, and now the peak energy demand can be shifted to the chosen base
machine speed. Thus, the converter capacitance can be designed according to:
symmetrical around the nominal operating point. In practice due to the second harmonic
in the arm power fluctuation, (5-40), the nominal capacitor will be in the range defined in
(5-31), but with an error from the desired average voltage VC. This error has been shown
144
General design for an n-module BMCI and experimental converter
between 0 and π/2. Thus, the method presented in this paper is valid for the design of the
capacitance.
routing section, and power circuit, consisting of the power sub-modules, floating
capacitors and power inductors. A high-level schematic of the converter power stage can
idc
SMx,a,1 iL,a,x
Cxa1
SMx,a,2 vxa
Cxa2
L =2Lm+Llk w
vin v
RBC vLL AC Machine
x=u
ix
SMx,b,1 L
Cxb1
SMx,b,2 vxb
Cxb2
iL,b,x x = u, v, w
The leg inductors are magnetically coupled and connected in such a way as to
maximize the inductance seen by the circulating current and minimize the converter’s
output inductance. The parameters of the components, and the test setup, are
145
Chapter 5
Symbol Value
Parameter
∆VL=100V)
Leakage inductance,
Llk 150µH
per winding
Sub-module
CSM 2.2mF
capacitance
Output line-to-line
Vline-to-line,peak 0 – 250V
voltage range
Device switching
fsw 5kHz
frequency
One of the design goals is to have flexibility in the converter so that only few hardware
changes would be needed if more functionality was required. Each sub-module assembly
consists of two parts – power PCB and control PCB. The power PCB carries the power
transistors, floating capacitor, gate driver circuits, isolated driver power supplies, and the
146
General design for an n-module BMCI and experimental converter
The sub-module was designed with 3 transistor legs, so that an energy source can be
connected to one leg, while the other two are part of the BMIC. This source can be either
The main transistor bridge, that is part of the BMCI, uses a quad IGBT module from
POWERX, supplied by Mitsubishi Corp., that is a partner on this project. While the
transistors might be considered oversized for a small laboratory prototype, they are
housed in a robust package and facilitate PCB and heat-sink mounting. A single sub-
147
Chapter 5
Figure 5-18 Single converter sub-module showing power PCB with IGBT power module and control card with
microcontroller. The MOSFET leg is not mounted on the PCB.
Each power transistor gate is controlled by an isolated gate driver, with a high current
output, part number Si8261-BAD. The chip uses a modulated RF-carrier for signal
transmission over the isolation barrier. The gate driver can source 1.8A peak, and sink 4A
peak current, which makes them well suited to driving large capacitive loads. The IGBT
module used is not a latest generation device and it has a relatively large total input
voltage. To avoid any Miller turn-on, the converter uses a bipolar gate drive, producing a
±15V gate-emitter signal. The gate driver includes the option for an external capacitor to
be connected across the gate terminals. This can be used if additional reduction of Miller
current effects. To protect the IGBTs, a pair of Zener diodes is also connected in parallel
The gate driver’s internal MOS transistors have typical internal resistance of RL = 0.8Ω
for sinking, and RH = 2.6Ω for sourcing, current. To ensure reliable operation the peak
148
General design for an n-module BMCI and experimental converter
output current of the gate driver must be limited to the rated value. This can be done by
RH Rg2 D
Vg+=+15V
Q
Vg-=-15V
RL Rg1
The peak turn-on current is determined by selecting the value of Rg1. With a bipolar
gate drive, the IGBT input capacitance will be charged from -15V to +15, making the initial
Vg + − Vg − 30V (5-46)
Rg1 − RH = − 2.6 = 14.066
I H , peak 1.8 A
During turn-off the gate driver has to sink both the gate current as well as the Miller
capacitor current that is caused by the high dv/dt seen by the gate-collector junction. To
fully utilise the gate driver’s sinking current capability, the turn-off gate resistance can be
lowered by introducing Rg2 and a series diode. This makes the turn-off gate resistance
Vg + − Vg − 30V (5-47)
Rg1 Rg 2 − RL = − 0.8 = 6.7
I L , peak 4A
By choosing Rg1 and Rg2 both equal to 15Ω the turn-on and turn-off currents are
limited to 1.705A and 3.614A, respectively. The schematic of the gate driver circuit used
149
Chapter 5
Figure 5-20 Isolated bipolar gate driver with separately isolated power supply
The BMCI prototype was designed with future work in mind. One of these project uses
a supercapacitor energy storage element, interfaced via a boost converter. When in use
this boost converter must regulate the sub-module capacitor voltage, monitor
supercapacitor voltage, and introduce limits to the energy storage element’s current. This
functionality requires two voltage sensors and one current sensor. To reduce the number
of signals measured by the high-level master controller, which would require isolation, a
control PCB card was designed. This card also has an isolation barrier, realised by using a
4-channel digital isolator, which has a much lower cost than an isolated voltage or current
transducer. The control card is fitted with a dsPIC33EP512GM604 digital signal controller
analogue inputs.
150
General design for an n-module BMCI and experimental converter
The 4 digital isolator channels are connected to 4 PWM outputs of the DSC and, when
the control card is inserted in the power board, to the isolated gate driver input, as shown
in Figure 5-21:
+3V3
Isolated
Rdr1
Digital gate driver
isolator
Master
Controller +3V3
Rdr2
DSC
CMOS
output
a.) b.)
Figure 5-21 Digital isolators a.) and drive of gate driver input b.)
The DSC output can be configured either as a normal CMOS output, an open-drain
output, or disabled. When configured as a normal output, the DSC controls the gate driver,
with its input current limited by Rdr2. When configured as an open drain, the top transistor
in the output stage is disabled, and Rdr1 functions as a pull-up resistor. In this state the DSC
output functions as an enable signal. Both modes can be used to bypass the master
controller output in case of a fault, e.g. output short circuit or sub-module overvoltage.
difference amplifier configuration. This circuit measures the voltage across a 5mΩ current
sensing resistor, vRsense, and produces a voltage vADC, centred around +1.5V. The transfer
151
Chapter 5
R2 (5-48)
v ADC = 1.5V + v Rsense
R1
To reduce the effects of the parasitic inductance of the current sensing resistor, and to
differential-mode filter. The circuit of the amplifier can be seen in Figure 5-22:
Figure 5-22 Difference amplifier circuit with current sensing resistor R17
The control system requires 6 current measurements – one for each arm. The
laboratory prototype uses isolated closed-loop current transducers, part number LA25-
NP. Each sensor is configured with 3 primary turns, giving a measurement range of ±12A.
The sensor functions as an ideal current transformer, with dc to 100kHz bandwidth, and
transfer ratio of 3:1000. The sensor is powered by a ±15V isolated power supply and can
produce an output voltage in the range of ±6.48V when loaded by a 180Ω load resistor.
152
General design for an n-module BMCI and experimental converter
+15V
Controller
Master
-15V
+15V
-15V
Figure 5-23 Current sensor design and placement with leg inductor
To manage all the connections between the master controller, the 12 sub-modules, and
other sensors, 2 routing PCBs were designed. Each routing card connects to two OPAL-RT
digital output modules, and one analogue input module, using a 38-pin ribbon cable for
each connection. These connections are routed to 6 sub-modules PWM signals and voltage
measurements, and up to two sensor cards, allowing for 4 sensor signals per routing card.
The converter only uses 6 current transducers, which makes one sensor card slot
153
Chapter 5
Figure 5-24 Connections between the power stages and the controller. Dashed connections are virtual.
simulator. This computer has 4 blocks that provide both digital and analogue
inputs/outputs. The high-speed controllers are implemented on an FPGA chip inside the
OP5600, while user controls are accessible through a desktop computer running
Simulink/RT-LAB. A diagram showing the connections between the power stage and the
The converter power stages, routing PCBs, and OPAL-RT controller are all housed in a
standard 19” rack and the three-phase output terminals are connected to either a 3-phase
5.3. SUMMARY
This chapter covered the design of a BMCI including the choice of number of sub-
modules n, calculation of total leg inductance and sub-module floating capacitance. The
154
General design for an n-module BMCI and experimental converter
number of sub-modules has a direct effect on the inverter ac-side voltage THD, with
higher values of n giving lower distortion. When operating in boost mode (ddc,x<0.5), with
maximum modulation index, the BMCI performs better than a conventional buck MCC, as
the converter can produce more discrete voltage steps, for the same number of sub-
modules.
The type of modulation scheme also has an effect, and PSC and PDC modulation are
examined. The main difference is that PDC modulation gives more consistent THD over a
range of ddc,x values, while PSC varies between the boundary points in Table 5-1.
Higher values of n increase the voltage steps in the output, reducing dV/dt, common
mode currents, and THD of voltage at motor terminals. This in turn has a beneficial effect
on motor efficiency, as iron losses are reduced. The size of the leg inductance also reduces,
while total arm capacitance remains constant for a given power level and base speed.
155
Chapter 5
156
Simulation of a BMCI for a hydrogen tram
This chapter shows the application of a BMCI in a light rail vehicle with a Hydrogen
Proton Exchange Membrane (PEM) fuel cell. Sub-module capacitance and leg inductance
values have been calculated using the methods described in Chapter 5 and the converter
solution using a voltage-source inverter with a cascaded boost converter (BVSI). The
criteria used are energy stored in capacitors and power inductors, total semiconductor
power rating required, converter efficiency, and total harmonic distortion. Catalogue
capacitor and inductor energy densities and specific energies are used to estimate the
have become a viable technology for the transport industry. This type of fuel cell also
benefits from temperature operation below 100°C. The BMCI proposed in this work is a
higher ac voltage can be useful for drives powered by a hydrogen fuel cell as PEM fuel cells
are characterised by a relatively high output resistance, which causes the terminal voltage
to vary over a wide range [155]. The BMCI can be used to compensate for this reduction
in stack output voltage and to interface the fuel cell directly to the traction motor, thus
157
Chapter 6
Most trams are powered by an overhead line, but the high cost of extending electrified
lines has created a niche for autonomous vehicles that carry the power source on board.
This has created a potential market for hydrogen fuel cell trams. The simulated BMCI has
been designed for a vehicle based on a CAF Urbos 3 tram [156], comprising of 7 articulated
120
Tractive effort (kN)
100
80
60
40
20
0
0 20 40 60 80
Velocity (kph)
The vehicle data includes the train resistance, according to the Davis formula [157],
with parameters for rolling resistance A, speed dependent resistance B, and aerodynamic
resistance C.
158
Simulation of a BMCI for a hydrogen tram
PMOT 12 x 80kW
Traction motors
resistance
efficiency
The exact details of the traction motors used in an Urbos 3 tram are not published and
are extrapolated from dc-link voltage and power rating. The synchronous frequency, slip,
power factor, and number of poles are chosen based on published data from traction
motor manufacturers [35], [158]. The Urbos 3 has 12 asynchronous ac motors, each rated
at 80kW continuous power. Standard practice in railway vehicles is to have one traction
inverter feeding two asynchronous motors, and for the simulation two machines are
lumped into one with twice the power rating with details listed in Table 6-2.
159
Chapter 6
532 VRMS
Nominal Voltage
Torque 772 Nm
Slip 1%
The vehicle mechanical model can now be constructed, using Table 6-1 and Table 6-2,
with a diagram shown in Figure 6-3:
C u2
+ B
+ A 0
TM vtram
3 kTE + 1/m
ωrotor kω
The model input is motor torque, TM, which is converted to tractive effort, TE, using
the coefficients kTE and kω. The rated speed is inferred from Figure 6-2 as 30 kph, with
160
Simulation of a BMCI for a hydrogen tram
30kph (6-2)
k =
207.34rads −1
The PEM fuel cell consists of seven basic elements, as illustrated in Figure 6-4. The fuel
cell generates electrical power from the chemical reaction of hydrogen and oxygen by
“routing” hydrogen electrons via the electric circuit connected to its output, while the
Without a load connected to the output, the charge build-up between the cathode and
the anode produces an ideal open-circuit voltage E0, that is described by [155]:
RTFC PH 2 PO2
0.5
E = 1.23 − k E (TFC
0
− 298.15) + , (6-3)
2 F 100 o1.5
, where TFC is the absolute fuel cell temperature in Kelvin, PH2 is the hydrogen absolute
pressure in kPa. PO2 is the oxygen absolute pressure, also in kPa, kE is temperature
161
Chapter 6
When the fuel cell is loaded, the terminal voltage drops due to three main mechanisms:
activation loss, ohmic loss, and concentration loss [159], [130], [160]. The resultant fuel
(6-4)
VFC = E 0 − (RAct − R − RConc )I FC ,
, where VFC is the fuel cell terminal voltage, RAct is the activation loss resistance, RΩ is the
ohmic resistance, RConc is the concentration resistance, and IFC is the fuel cell output
current.
At high currents, the fuel cell concentration resistance increases exponentially due to
mass transport, with terminal voltage eventually reaching 0. In a vehicle application, the
fuel cell stack rated power would be somewhere below the peak power, where output
voltage drop is dominated by activation losses, as increasing the current any further risks
The detailed model of the fuel cell stack requires parameters such as temperature
during operation, hydrogen and air pressures, and other data specific to the
manufacturer. This information is not readily available and instead a simplified model is
Cac t
RΩ R ac t
E0
VFC VFC
Since the fuel cell would be operated only in its stable region, non-linearities at high
output current can be ignored. Below that region the fuel cell can be modelled as a voltage
162
Simulation of a BMCI for a hydrogen tram
source, with a series resistance and a parallel capacitance, to account for the transient
response. It should be noted that the simplified circuit model does not consider the
chemical and mechanical dynamics of the fuel cell power and assumes static resistances.
The simulated fuel cell stack is based on 2 Hydrogenics HD90 [161] units connected in
parallel, that can supply up to 198kW of continuous power. The actual fuel cell parameters
are unknown and are extrapolated from data found in literature [162] and [163], shown
in Table 6-3. The ohmic and anode resistances are chosen to be 10% of the total series
resistance each. The capacitance of the cell is related to the fuel cell active area and is thus
Symbol Value
Parameter
A plot of the static V-I characteristics, at dc, of the simulated stack are shown in Figure
6-6:
163
Chapter 6
drive, producing the same peak line-to-line motor voltage, but drawing power from a
hydrogen fuel cell. During braking the energy, recovered by the converter, is dissipated
into a braking chopper resistor. This design uses no energy storage and the fuel cell is
designed to supply the peak power demand of the system, including auxiliary loads.
H2 O2
Figure 6-7 Conventional motor drive (left) and proposed BMCI fuel cell drive (right)
164
Simulation of a BMCI for a hydrogen tram
The numerical simulation model uses the continuous time equations derived in
Chapter 3, that describe the system. The converter inputs and outputs are maintained by
the control systems, described in Chapter 4, with the addition of a speed controller to
simulate a tram driver. The BMCI is designed using the design methodology in Chapter 5.
Value
Parameter
frequency
The switches in the simulated model have the same series resistance as a
cruising – braking traction cycle. The simulated induction motor drive accelerates the
vehicle up to 50kph, which is a typical maximum speed for an urban tram, and brakes
back to a standstill. During this cycle all power is provided by the hydrogen fuel cell stack,
165
Chapter 6
The BMCI – fuel cell drive, with all control systems operating, is simulated for a full
traction cycle, displayed in Figure 6-8. The figure shows the speed trajectory of the BMCI
powered tram, as well as fuel cell stack terminal voltage and output current, motor torque,
During the initial acceleration, the BMCI is operating in voltage buck mode. At this
point the fuel cell stack voltage is at its highest, while the output voltage is at its lowest.
compensation, required to operate with low output frequency. The current ripple caused
by small unbalances in the compensation currents between phase legs can be seen to
The converter’s boost ratio increases in a non-linear fashion, as the fuel cell voltage
drops while output power increases. The highest boost ratio achieved is about 3.125, and
166
Simulation of a BMCI for a hydrogen tram
it occurs when the converter is operating with a constant power output. As input power
reduces, fuel cell stack voltage increases, reducing the boost requirement. While the
converter can be controlled with variable sub-module capacitor voltage, in the scheme
presented here it is kept constant to reduce capacitor voltage ripple amplitude. In Figure
6-9 the average arm sub-module voltages, motor current, and leg circulating currents are
shown.
Figure 6-9 Feedback values for the BMCI during a traction cycle
At start up the converter arm currents have a high amplitude, due to the injected
is disabled, and the second harmonic suppression controller is engaged, which accounts
for the disturbance to the capacitor voltage ripple. However, the converter input current
in Figure 6-8, has a much lower current ripple, as the large oscillations are circulated in
167
Chapter 6
During the whole traction cycle the peak-to-peak sub-module capacitor voltage ripple
When the vehicle is moving in the low-frequency region, which for the studied example
is for an output frequency below 28Hz, the BMCI inserts in each leg a differential voltage
and injects a common mode current at the same frequency. The inserted voltage must be
the same for all converter legs so as not to appear across the fuel cell stack or the motor
terminals.
By zooming in at the start of the cycle, shown in Figure 6-1, converter terminal voltages
and currents can be better examined in Figure 6-10. It can be observed that the motor
voltage has relatively low quality, as the voltage steps consist of more than one level.
Figure 6-10 Converter voltages and currents at start of acceleration showing low distortion machine current
168
Simulation of a BMCI for a hydrogen tram
Before start-up, at standstill, the inverter is maintaining constant field current to avoid
charging the magnetizing inductance during the traction cycle. When a torque command
is applied, the inverter produces slip, at which point the vehicle starts accelerating. The
PWM waveform pulses cannot be distinguished due to the high frequency ratio fPWM/ffield.
A further magnification of the converter operation around t=0.95s shows the dc-side
current ripple and the shape of the ac-side voltage pulses, as Figure 6-11 shows.
Figure 6-11 Zoomed in capture at t=0.975s, showing dc current ripple caused by the low-frequency arm-balancing
currents; motor currents only show small spikes from the arm voltage step switching
The BMCI input current contains both switching ripple and the injected compensation
current at 400Hz, with the latter becoming more visible as frequency increases. This is
due to the decreasing ratio of converter output frequency and compensation frequency.
169
Chapter 6
The goal of the low-frequency compensation scheme is to limit the voltage ripple
across the sub-module capacitor for each arm. Figure 6-12 shows how the common mode
current is injected in each arm current, effectively chopping the phase current. The
component that accounts for energy transfer between the top and bottom arms appears
Figure 6-12 Arm capacitor voltage, arm currents, and circulating current at low machine frequency
The demand for the compensating current is highest at start-up, as at that point there
is little power drawn from the fuel cell, its voltage being close to the maximum value, and
the voltage ripple is at a low frequency. The simulation shows that even in demanding
overshooting the design’s maximum peak-to-peak value or the peak arm current value.
170
Simulation of a BMCI for a hydrogen tram
In the constant torque region, the converter initially bucks and then boosts the ac-side
With the peak line voltage less than the fuel cell terminal voltage, the modulation index
M is less than the dc duty cycle ddc, and the number of effective output levels is low,
reducing the motor voltage distortion. The converter voltages and currents, at the time of
Figure 6-13 Converter in the constant torque region, with ac-side voltage buck
Before the point of transition, the BMCI dc-side current has a peak-to-peak ripple of
80A at 400Hz. When the output frequency is high enough and the ripple compensation is
no longer required, the converter returns to its normal operations. The initial disturbance
is due to the activation of the circulating current harmonic suppression controller. Also, it
can be noticed that the phase current has much lower current ripple as the phase voltages
171
Chapter 6
no longer have the higher voltage steps caused by the low-frequency ripple compensation
scheme.
Referring to Figure 6-8, the converter enters boost operation at roughly t=8s. After this
point the converter ac-side voltage gains additional steps and the waveform quality
In the constant power region, the BMCI is operating at fixed boost ratio, with constant
ac-side voltage amplitude, but with increasing machine slip. The ac-side voltage waveform
has the maximum number of levels, for the appropriate dc duty cycle. A zoomed-in view
of the waveform from the converter traction cycle is displayed in Figure 6-15.
172
Simulation of a BMCI for a hydrogen tram
Figure 6-15 BMCI terminal voltages and currents at constant output power
At constant power the fuel cell is producing 220V, with an output current of 700A, and
the dc duty cycle ddc ≈ 0.191. The voltage of a leg inductor, and its frequency spectrum are
173
Chapter 6
Figure 6-16 Leg inductor voltage and its frequency spectrum during vehicle acceleration; peak component is at the
arm switching frequency
The figure shows that the leg inductor voltage switching step is only one sub-module
voltage, with only a few cycles of 2-step switching. The frequency spectrum shows that
most of the harmonic energy is around 2n times the device switching frequency fsw, and
During regenerative braking the BMCI works at the lowest boost factor, as it draws no
power from the fuel cell stack while energy is dissipated by the braking resistor. This
change is that during braking ddc remains constant and is different form the value during
174
Simulation of a BMCI for a hydrogen tram
acceleration. This change in dc duty cycle affects both input current ripple and output THD
Figure 6-17 Converter currents and voltage at start of braking; ac-side voltage waveform changes as converter is
regenerating power and dc duty-cycle steady-state value is different
After a reverse torque step at t=32s the converter input current is reversed to -390A
ddc=0.315. The spectrum of the voltage across one phase inductor is shown in Figure 6-18.
175
Chapter 6
Figure 6-18 Phase leg voltage and its FFT spectrum during vehicle braking
Similar to the acceleration phase, the inductor voltage waveform is also switching with
1 voltage step and a few pulses of 2-step. The frequency spectrum, however, is different,
with lower energy around the 2nfsw point (4th harmonic), and more energy shifted to 8th
harmonic. With the converter ac-side voltage having roughly the same amplitude, the
study has been undertaken, based in the same fuel cell tram drive as in the rest of this
chapter. It compares the BMCI with the state-of-the-art topology of a boost inverter
(BVSI), composed of an interleaved boost dc-dc converter and a traditional VSI. Both
176
Simulation of a BMCI for a hydrogen tram
topologies are designed for the same traction motor, and fuel cell parameters, as listed in
Table 6-2 and Table 6-3. The comparison has been based on total semiconductor apparent
power, mass and volume of energy storage components, efficiency and weighted total
harmonic distortion (WTHD), as these parameters are important for a traction system.
The passive components for the two systems have been designed based on the
optimized for a specific application, the power density and weight of the power
electronics is not evaluated. Gate driver losses are also ignored, as they are part of the
system realization and can be difficult to predict. However, they are accounted for in the
In part 6.3.5 system efficiency and weighted total harmonic distortion (WTHD) are
evaluated. To achieve static conditions, the simulated load inertia is set to be 1000-times
higher, resulting in constant machine speed for the duration of the numerical simulations.
2
Vnharm
nharm = 2,3,... nharm
(6-5)
WTHD =
V fund
, where Vfund is the amplitude of the fundamental and nharm is the number of the harmonic,
6.3.1.1. BMCI
The worst-case arm current occurs at rated power, and unlike voltage-source
inverters, the BMCI switches do not need to be rated for the peak machine current at
standstill – the dc-side current at that point is very small. Instead, the arm RMS current at
2
n
I v (6-6)
Sarm = I arm,rms VC ,i = I dc , x 2
+ ac , x n VC + C
i =1 2 2
VC=300, ΔvC=60):
n
Sarm = I arm,rms VC ,i = 288.5 2 330 = 190.4kVA (6-8)
i =1
6.3.1.2. BVSI
The boost converter devices must be rated for the full-power fuel cell current. The
The voltage-source inverter must be rated for the peak machine current, as the
system must be able to operate down to 0Hz motor frequency, with the full rated stator
current.
An important note is that the BVSI Sboost is for a diode and transistor combination, and
178
Simulation of a BMCI for a hydrogen tram
The BMCI capacitance is designed for 20% voltage ripple, while the boost-voltage
source inverter dc-link capacitor is designed for 1% voltage ripple. This is mainly a
function of the required capacitor ripple current and less so of dc-link voltage stability
requirement.
6.3.2.1. BMCI
VC ,nom 2 (6-13)
WC , BMCI = 6nC = 5.4kJ
2
converter, with the largest oscillating power component being at the fundamental
machine frequency.
6.3.2.2. BVSI
As the dc-link capacitor is common to all 3-phases of the inverter, and the two phases
of the boost-converter, the required capacitance is much lower. In addition, the boost
converter PWM carrier can be shifted, so that it is complementary with the inverter dc-
side current pulses. This configuration reduces the capacitor RMS current.
179
Chapter 6
3.00E+05
electrolytic capacitors, where increasing the voltage by a factor of 2.14 gives an increase
of 20%, which means a single large capacitor will be a denser solution, than multiple
lower-voltage capacitors.
The specific energy of the same capacitors is shown in Figure 6-20. The data for some
250
Specific energy (J/kg)
API
200
Vishay
150 SBE 700A105
SBE 700A136
100
Ducati DC85C 550V
50
Ducati DC85C 700V
0 Ducati DC85C 1.5kV
0.0000 0.0020 0.0040 0.0060 0.0080 0.0100 0.0120
Capacitance (F)
180
Simulation of a BMCI for a hydrogen tram
Using the DC85C series capacitors, from Figure 6-19, the total capacitor bank weight
and volume are estimated for each converter. These capacitors were chosen, as they
represent typical parts used for energy storage and filtering applications in power
converters: they have high ac-ripple current ratings and are characterised for the working
i. BMCI
The BMCI requires 12 capacitors with total energy stored of 5.4kJ. Using the energy
density of the 550V DC85C capacitors, the total capacitor bank volume is:
5.4kJ
Vol C , BMCI = = 0.027m3 = 27l (6-15)
2.0110 J
5
3
m
The total capacitor mass is calculated using data from Figure 6-20:
5.4kJ
massC , BMCI = = 27.6kg (6-16)
195.56 J
kg
ii. BVSI
The BVSI uses only one capacitor, but it requires lower voltage ripple, and a higher
352 J
Vol C , BVSI = = 0.0011m3 = 1.1l (6-17)
3.2 105 J 3
m
181
Chapter 6
352 J
massC , BVSI = = 1.28kg (6-18)
274.2 J
kg
Compared to the BMCI, the BVSI capacitor bank will be roughly 6 times smaller and 5
times lighter.
The governing parameter for leg or phase inductance are the peak-to-peak ripple
currents ΔiL,leg and ΔiL,phase, for the BMCI and BVSI respectively. They are designed to be
6.3.3.1. BMCI
Using the constraints above and the tram data from Table 6-1, the dc-current, phase
From which the peak energy stored in each of the three leg inductors is:
i
2
Lleg I dc , x + dc , x (6-20)
WL , x =3 2
= 7.03J
2
The BMCI current ripple is not constant over time, and the peak value is maintained
only for a certain portion of the output waveform cycle. Thus, inductance can be optimized
182
Simulation of a BMCI for a hydrogen tram
6.3.3.2. BVSI
Repeating the design procedure for the BVSI, and with the same constraints as for the
BVSI, the dc-side current, phase dc current, current ripple and phase inductance are:
From which the peak energy stored in each of the phase inductors is:
i
2
Lphase I dc , phase + dc , phase (6-23)
WL , x =2 2
= 96.9 J
2
The interleaved boost converter reduces the total dc current ripple to 64% of the
amplitude of each phase, compared to the BMCI, where it was measured as twice the leg
value.
Power inductors can occupy half of the converter volume and can account for most of
the system weight and cost [41], [76], [169]. This is reflected in their poor energy density
and [171], is shown in Figure 6-21. The inductors were chosen to have a low inductance
183
Chapter 6
The specific energy, of the inductors in Figure 6-21, is shown in Figure 6-22:
3.50E-01
3.00E-01 High Ripple
2.50E-01 HCS-301M-1000A
2.00E-01 19026L
1.50E-01 CWS-1EE-12721
1.00E-01
HCS-121M450A
5.00E-02
MegaFlux
0.00E+00
0.0000 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006
Inductance (H)
To compare the two systems, in terms of total weight and volume of all inductors, the
calculations are done using data from Figure 6-21 and Figure 6-22.
184
Simulation of a BMCI for a hydrogen tram
i. BMCI
The BMCI uses 3 inductors with total energy of 7.03J. Using the CWS-1EE-12721
7.03J
Vol L, BMCI = = 0.0184m3 = 18.4 l (6-25)
0.382 10 J
3
m3
7.03J
massL, BMCI = = 41.59kg (6-26)
0.169 J
kg
ii. BVSI
The cascaded boost converter and voltage source inverter require total inductor
energy storage of 42.54J, which results in total volume and weight of:
96.9 J
Vol L, BVSI = = 0.0541m3 = 54.1l (6-27)
1.79 103 J
m3
96.9 J
massL, BMCI = = 259.1kg (6-28)
0.374 J
kg
When inductors are utilized for power converters, their design is a balancing act
between ac and dc winding losses, mainly due to the proximity effect losses. To evaluate
the dc conduction losses, for the power inductors surveyed, the ESR of each part is plotted
in Figure 6-23:
185
Chapter 6
2.00E-04 HCS-121M450A
MegaFlux
0.00E+00
0.0000 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006
Inductance (H)
These values are used to extrapolate the inductor ESR for the efficiency calculations in
part 6.3.5.
The efficiency of each converter has been calculated by simulating the converter and
measuring the dc-side and ac-side voltages and currents. As the 3-phase voltage has a
switching component, a low-pass filter is used to calculate the average value of the
efficiency.
LPF
ηconv (%)
vdc vac
idc iac
Converter
AC
Under Motor
VFC Test
186
Simulation of a BMCI for a hydrogen tram
6.3.5.1. BMCI
i. Designed BMCI
The BMCI design values are summarised in Table 6-5, and the converter schematic is
shown in Figure 6-25. The inductor properties are extrapolated from the values for part
CWS-1EE-12721:
Converter legs/phases 3
187
Chapter 6
SMa,x,1
SMa,x,n x = u, v, w
300V
w
Lleg
M
x=u
SMa,x,1
SMa,x,n
ib,x
The BMCI numerical simulation model considers conduction losses in the power
switching loss. Power inductor ac-resistance and core losses are not accounted for, as
there is no data available from the manufacturers surveyed. Transistor and diode
conduction losses are modelled as an ideal voltage source in series with a parasitic
resistance, while the switching losses are modelled as a controlled current source, in
parallel with the sub-module capacitor. A simplified diagram is shown in Figure 6-26:
188
Simulation of a BMCI for a hydrogen tram
2vSi
1-dx,a,i dx,a,i
rC
ix,a
vout,SM,x,a,i
vSM
C dx,a,i
2rSi
1-dx,a,i
n
vout,SM,x,a,i ix,a
iloss,x,a/b,i C i=1 Vdc
vSi=f(iSi)
iSi rL 2
C rSi
= iSi L vx
E E
vout,SM,i
>0
ix,a ZOH 1/VcrefTstep iloss,x,a/b,i
<0
T=Tstep
aon+arr
bon+brr
aoff
boff
Figure 6-26 Lossy BMCI model for a top arm with IGBT switches.
pulses, scaled by the switch current, and triggered by a rising or a falling edge of the sub-
module ac-side voltage vout,SM,x,a/b,i. The transistor and diode conduction loss are accounted
, where iSi is the semiconductor switch current, and Vsat is the junction saturation
voltage, which is assumed to be equal for the IGBT and the diode.
To simplify the simulation of the switching losses, instead of a quadratic fit the
189
Chapter 6
, where Eon is the turn-on energy, Eoff is the turn-off energy, and Err is the diode reverse
recovery energy, which can be found in most device datasheets. The coefficients aon, arr
and aoff are the intersection between the loss curve and the y-axis, while bon, brr and boff
are the proportional coefficient that account for the losses’ dependency on switch
current. VCE,ref and ICE,ref are the collector-emitter voltage and collector current used in
the manufacturer’s tests, and Tstep is the discrete time step used in the numerical
simulation.
The simulated BMCI efficiency was measured to be ηBMCI= 92.5%. Of the pproximately
13kW losses, 11.1kW are semiconductor conduction losses, 900W are switching losses,
190
Simulation of a BMCI for a hydrogen tram
The WTHD level was simulated at rated power with frequency ratio of 36 for both the
BMCI converter and the BVSI system, resulting in electrical frequency of roughly 111Hz.
The BMCI line-to-line voltages, and phase currents, are show in Figure 6-27:
6.3.5.2. BVSI
i. Designed BVSI
The designed BVSI parameters are summarized in Table 6-6. The inductor is
191
Chapter 6
Converter legs/phases 2
The simulated BVSI losses are simulated the same way as in the BMCI case, with the
switching loss current being drawn from the dc-link reservoir capacitor, shown in Figure
6-28:
192
Simulation of a BMCI for a hydrogen tram
0.5rL rC
800V
AC
Lphase Motor
iloss Cdc
VFC
C
Lpha se vSi=f(iSi)
iSi
C rSi
vu = iSi
vv
vw E E
va
vb
>0
iu ZOH 1/VcrefTstep iloss
iv <0
T=Tstep
iw
ia aon+arr
ib bon+brr
aoff
boff
5.2kW loss breaks down into 1.1kW of switching losses, 3kW of inductor loss, and the
The 2-level voltage source inverter was measured to be WTHDBVSI = 1.5%, which
results in higher machine ripple current. The BVSI ac-side voltages and currents are
193
Chapter 6
The results of the comparison show that the BMCI’s inductors are much smaller than
the BVSI ones, even though the required capacitance is much larger. Therefore, the BMCI
system total energy storage components are smaller and considerably lighter, fulfilling
one of the most stringent requirement of traction systems. The main drawback of the
BMCI is its lower efficiency at full load, resulting in a larger cooling system that would
partly reduce the gains of smaller energy storage components. The much lower machine
WTHD achieved with the BMCI is expected to improve machine efficiency. However more
The resultant parameters of the two converters are summarised in Table 6-7.
194
Simulation of a BMCI for a hydrogen tram
6.4. SUMMARY
This chapter has presented a numerical model for a fuel cell tram with a BMCI
induction motor drive. The converter is put through a traction cycle to demonstrate
operation in all modes required for an electric vehicle: constant torque, constant power,
and regenerative braking. The converter operates at output frequencies from 0 to beyond
rated speed, with a constant current amplitude in the linear power range, and with
a BVSI system is also designed and simulated. The two converters are then compared in
terms of energy storage volume and weight, total switch apparent power needed,
195
Chapter 6
efficiency, and harmonic distortion. The BMCI is shown to have the advantage of
significant reduction in energy storage component weight and electric motor current
WTHD, which can reduce the chance of machine insulation failure and increase its
efficiency. The main disadvantage is the lower peak efficiency, although a more accurate
analysis would be needed to evaluate the impact on the total energy lost in a practical
traction cycle and the actual average thermal load of the cooling system.
196
Experimental validation
This chapter presents experimental data from the prototype BMCI. Initial tests have
been carried out without feedback of currents or voltages, to validate the basic working
principle and verify the boost characteristics. The converter has been then connected to
an induction motor to demonstrate operations over the full speed range required by an
electric vehicle, with the converter’s ac-side RMS voltage higher than the dc source. The
idc
SMx,a,1 iL,a,x
Cxa1
SMx,a,2 vxa
Cxa2
L =2Lm+Llk w
vin v
RBC vLL AC Machine
vu x=u
ix
SMx,b,1 L
Cxb1
SMx,b,2 vxb
Cxb2
iL,b,x x = u, v, w
The dc-source voltage vdc is varied between different tests to avoid the bench power
supply going into power limitation mode. Open-loop tests examine the maximum boost
characteristics and at very low dc duty cycle the dc-side current is limited by the power
supply current limit, output load, and converter parasitic resistances. When the converter
is used with the induction motor, the dc-side voltage can be raised, as the maximum dc
source power can be controlled by adjusting the machine base frequency for field
197
Chapter 7
weakening mode. A photograph of the hardware setup, showing the converter, OPAL-RT
Figure 7-2 Final setup of the converter prototype with a variac load, induction motor load, and dc power supply.
198
Experimental validation
with open loop control – a fixed duty cycle ddc, fixed output frequency fout, and fixed
amplitude of the ac duty cycle dac. The only closed-loop controller that a modular cascaded
converter requires is the arm-balancing algorithm that ensures all capacitors, within the
achieve a variable load, the resistors are connected to the converter by a 3-phase variable
side voltage ratio, Gvariac can be varied between 0 and 100%. At 0% the only load seen by
load. Maximum load is achieved when the autotransformer output is set for 100%, in
which case the per-phase resistance seen by the converter is equal to a third of the power
resistors, or 1.867Ω.
For this part of the experiment, the converter waveforms are examined in 3 cases of
ddc: 0.3, 0.5, and 0.1. For each case the load is adjusted to achieve output current amplitude
of 2A peak. To avoid core saturation of the variac, the output frequency fout is set to
208.33Hz. The value was chosen to get an integer number of PWM pulses per cycle. With
a switching frequency fcarrier of 5 kHz per transistor, the frequency ratio, fcarrier/fout is equal
to 24. However, the effective switching frequency is 4 times higher, as the sub-module
199
Chapter 7
Figure 7-3 Converter phase voltages, referenced to power supply 0V, and output currents
Figure 7-3 shows the converter running with ddc=0.3, dac=0.7×sin(2π×208.33×t), and
Gvariac=0.325. The variac leakage inductance, together with the changed ratio of passive
resistance to inductance, can clearly be seen, as the switching harmonics are not visible
and the current waveform, and the currents have considerable displacement angle. As the
dc-side power supply is fixed at 30V the BMCI is operating with a low voltage-boost factor
Gv=Vll,RMS/Vin=1.2. The primary reflected load resistance is calculated using (7-1), and is
equal to 16.25Ω.
Rload (7-1)
Rload '=
Gvariac
The oscilloscope capture shows the output phase voltages Vu, and Vv. The voltages are
offset by one half of the dc-side voltage, 15V, and a total of 7 levels can be seen, with a step
of about 12V. The arm voltages of phase u, as well as the arm currents, are shown in Figure
7-4 and the average capacitor voltage was measured to be about 22V, by compensating
200
Experimental validation
for the IGBT saturation voltage drop. Each arm can produce 4 switching levels at
maximum modulation index. These results conform to the predictions made in Chapter 5.
Figure 7-4 Arm voltages and currents at ddc=0.3, with maximum current ripple
harmonic and a second harmonic that is clearly visible in the circulating current.
The peak current ripple occurs when edges of the top and bottom arm switching
waveforms align. At these intersections the leg inductor waveform switches at frequency
fsw,eff=n×2×fcarrier and has a voltage step equal to 2×VC. This can be observed in Figure 7-4,
at every crossing and peak of an arm current waveform. It should be noted that the arm
current peaks do not align with the peak voltage, as the load is inductive.
201
Chapter 7
The next step looks at the converter operating at dac,max=ddc. In this condition, the BMCI
operates as a conventional buck MCC. As the carrier shift is 0 between the two arms, the
pulses at the phase terminal align and the voltage step is equal to VC, with operation
equivalent to a conventional 3-level inverter. The voltage waveforms and ac-side currents
Figure 7-5 Converter phase voltages, referenced to power supply 0V, and output currents, at ddc=0.5
The low number of voltage levels causes a higher ac-side current ripple, compared to
the case of ddc=0.3. However, the dc-side current ripple is lower as every rising edge from
the top arm should be matched by a falling edge in the bottom arm, and vice versa. In
sub-modules will cause different instantaneous duty cycle for each module, to maintain
equal voltage. In addition, the top and bottom arm voltages are not equal at all times, and,
202
Experimental validation
even with a perfect edge match, there will be a small voltage applied across the inductor.
This can be observed in Figure 7-6, specifically between t=-3.5ms and t=-2.5ms, and every
The circulating current’s switching ripple is only visible at the peaks of the arm
currents. The voltage spikes in the inductor voltage waveform are due to the arm voltage
balancing controller. This effect depends on the direction of current flow, and the polarity
of the arm voltage. Investigation of this effect is left for future work as the pulses have
When the converter is pushed into a deeper boost, with ddc<1/2n, the converter
achieves the maximum possible number of output steps, as the modulation signal will
Figure 7-7 Converter phase voltages, referenced to power supply 0V, and output currents, at ddc=0.1
Under these conditions the BMCI produces phase waveform with 9 voltage steps of
½×Vc, but in most regions the voltage is switching between 3-states. The arm voltages
have 5 levels, cycling from -2VC to +2VC. The arm voltages, currents, and inductor voltage
are shown in Figure 7-8. The inductor waveform’s switching frequency is either 8fcarrier
or 4fcarrier, depending on the value of the modulated ac signal, with peak ripple current
occurring when a rising edge from the top arm coincides with a rising edge from the
bottom arm and the inductor voltage has a step of 2Vc. The other condition for high ripple
current is also visible, between t=-3.5ms and t=-3ms, where only one of the two switching
edges coincide, and the top and bottom arm pules are back-to-back. At this point the
204
Experimental validation
As the boost ratio is GV=3.833, the arm currents are always positive, and each arm
angle.
To investigate the maximum boost ratio of the converter the load resistance and dc-
side voltage are held constant, while the ddc is cycled from 0.075 to 0.6. The amplitude of
205
Chapter 7
Table 7-1 Experimental setup parameters for deriving maximum boost characteristics
Autotransformer
Gvariac 0.2
output ratio
Device switcihg
fcarrier 5 kHz
frequency
A sweep of ddc from 0.075 to 0.60 can be seen in Figure 7-9 through Figure 7-11,
demonstrating both ac-side voltage boost and buck. Maximum ac voltage is achieved with
ddc=0.075, i.e. the smaller the dc duty cycle, the higher the gain. The data shows the line-
to-line voltage Vll, dc voltage Vdc, and floating capacitor voltage Vcap.
206
Experimental validation
Vdc
Figure 7-9 Sub-module capacitor voltage, line-to-line voltage, and dc-side voltage for ddc=0.075 through 0.2
For values of ddc<0.2 the submodule capacitor voltage is larger than the dc-side voltage,
which is specific to the particular number of submodules per arm. With these low duty
cycle values, the modulated ac waveform’s zero-crossings are close to 0, which is one
point where minimum dc-side current ripple occurs. As discussed in Chapter 5, achieving
the minimum dc-side current ripple, and utilizing the maximum converter number of
output levels, are mutually exclusive conditions. It is expected for the number of discrete
levels to increase, as the duty cycle approaches 0.25. This can be observed in Figure 7-10.
It may appear that at ddc=0.3 the ac-side waveform is cleaner, but what may look like
207
Chapter 7
Vdc
Figure 7-10 Sub-module capacitor voltage, line-to-line voltage, and dc-side voltage for ddc=0.25 through 0.4
As the duty cycle approaches another point of minimum dc-side current ripple, ddc=0.5,
the number of ac-side voltage levels reduces, and the line-to-line voltage has more pulse
alignments, thus having less defined voltage levels. This is shown in Figure 7-11:
208
Experimental validation
Vdc
Figure 7-11 Sub-module capacitor voltage, line-to-line voltage, and dc-side voltage for ddc=0.45 through 0.6
The experimental results are compared to the large-signal model and ideal model in
Figure 7-12. They show good agreement with the analytical model.
Figure 7-12 Comparison between maximum ac-side voltage boost between hardware prototype and analytical model
209
Chapter 7
The voltage gain is lower than the ideal case, which is due to the voltage drops across
the IGBTs, the parasitic resistance of the capacitors, and the internal impedance of the
voltage source. With two arms connected in series and two devices per sub-module
always conducting at the same time, assuming vFWD=vIGBTsat, the dc-side voltage is reduced
by 5.6 V. In the experiment the main reason for keeping the dc-side voltage low is the
all equal. This condition is maintained by a set of 3 closed loop controllers that undertake
the following functions: sub-module balancing, arm balancing, and average leg capacitor
balancing. It is important for all of them to function under dynamic conditions and their
operations are examined using the traction cycle defined in the previous section.
At the end of the section the converter is tested at low frequency, demonstrating the
Spread in capacitance between different capacitors with the same part number can be
as high as 20%. To maintain the sub-module capacitor voltage below a maximum rated
value, the duty cycle of each floating capacitor must be slightly different. Figure 7-13
210
Experimental validation
Figure 7-13 Submodule voltages for the top and bottom arm of one converter leg.
The voltages of arm a have an offset of about 2 volts. This is due to offsets in the
internal voltage transducers as well as the probe used for the measurement.
The arm balancing controller needs to maintain the difference between the top and
bottom arm average capacitor voltages equal to 0. They can be observed, during a
211
Chapter 7
Figure 7-14Average top and bottom arm capacitor voltages during a transient
The offset between the two arms is small, but remains constant, as it is due to
measurement errors.
Each arm’s voltage regulator maintains the average voltage equal to the reference. This
method uses a single controller to maintain both the dc value, as well as achieve balancing
between the converter legs. Figure 7-15 shows capacitor voltage before and after a step
212
Experimental validation
Figure 7-15 Step response of capacitor voltage reference; average arm capacitor voltages (top) and dc-side and
phase currents (bottom)
The figure shows the effectiveness of the controller and that it is capable of balancing
the sub-module capacitor voltages during a transient. The modulated signal is kept
constant at m and the output current increases accordingly. During the initial response
the dc power supply goes into current limiting, causing a small oscillation at the start of
the transient.
When the converter output frequency falls below 5 Hz the capacitor ripple starts
increasing. Without additional measures the ripple can reach an infinitely high value at
0 Hz. As this makes the converter unstable, a capacitor voltage ripple compensator has
been introduced. This compensator aims at maintaining the top and bottom arms
balanced by inserting a common mode component at 100 Hz. Figure 7-16 shows the
transient of sub-module capacitor voltage, converter dc-side current, and phase u output
Dc-side
The 100 Hz ripple is clearly visible on the waveform of the capacitor voltage and the
ripple is less than 5V. The dc-side current also has a small 100 Hz component that has to
be provided by the dc source. In Figure 7-17 the arm currents can be seen, where the
214
Experimental validation
Arm currents
Figure 7-17 Sub-module voltage and arm currents with approximate fout=2Hz
operations. During these experiments, the average floating capacitor voltage has been
field-oriented control (FOC) system. Internal balancing controllers maintain the same
average capacitor voltages across all the sub-modules. At low frequency, the BMCI uses
The ratio of field current and torque current has been adjusted to achieve field
weakening at a base speed lower than the actual rated speed of the motor, so that the
converter boost ratio stays within the stable region. The test setup and induction motor
215
Chapter 7
Floating capacitor
VC* 80 V
voltage reference
Stator parasitic
Rs 8.63 Ω
resistance
Stator leakage
Ls 50.26 mH
inductance
torque limited to the motor’s rated value. The ratio of Id/Iq was adjusted to achieve
approximately 400 W maximum dc power, with acceleration time of less than 9 seconds.
To simulate a realistic electric vehicle, a traction cycle has been designed with
acceleration up to a target speed of 1200 rpm, cruising, and braking to standstill. During
216
Experimental validation
this cycle, the converter’s line-to-line voltage, dc voltage, phase current, and dc-side
current are monitored. The waveforms are displayed in Figure 7-18. The dc-side power
increases linearly during constant torque operations, while it remains roughly constant
during field weakening. The overshoot of the dc-side power is attributed to the delay of
When the converter starts braking the induction motor, the tachometer delay causes
the wrong electrical angle to be calculated. As the measured motor speed is not the true
speed, the actual machine slip is different from the calculated one. The motor voltage
amplitude has a smaller gradient, but there is an increase from the value at cruising
speed..
To gather all the measurements the experiment is ran in two parts – acceleration to a
constant speed, then braking from a constant speed. The results are then concatenated to
get a single traction trajectory. The effective converter switching frequency is in the range
of 20-40 kHz, and each experiment part lasts 10s; the voltage waveforms are low-pass
filtered to show only the fundamental harmonic. The voltage boost ratio, GV, is calculated
using a moving RMS window, whit a fixed window length, which causes artefacts, which
217
Chapter 7
During the initial acceleration, the drive is operating in constant torque mode, with the
output current having a constant amplitude. The ac-side voltage increases linearly, after
an initial step to overcome the stator resistance. When the machine enters field
weakening mode, the ac-side rms voltage remains constant until the torque reduces at
time=7s. The braking starts at time=10s, with a torque command step. At that point, the
converter starts regenerating the flywheel energy and dissipates it in the braking
chopper. This is apparent as the dc-side voltage raises to roughly 53V. At time=15s, the
drive can no longer regenerate any energy, and power from the dc supply is required to
brake the machine further. From the analysis of the traction cycle, the peak dc power is
Throughout the traction cycle the average capacitor voltage is 80V and is kept
constant, while the voltage ripple varies with the output frequency, as shown in Figure
7-19.
218
Experimental validation
Figure 7-19 Arms a and b capacitor voltages and currents during the traction cycle
At stand-still the converter output power is very low, and the finite precision of the
arm current measurement cause larger voltage errors in the steady-state. However, with
higher power throughput the balancing systems maintain equal voltages between sub-
The following sections examine the converter’s voltages and currents by zooming into
Figure 7-18. Special attention is payed to operations with output frequency below 5 Hz
7.3.1.1. Acceleration
Before acceleration begins the motor flux is maintained constant. This avoids the need
of excessive slip at start-up, which would be the case if the motor magnetizing inductance
is not charged. With a constant current id in the stationary state, the converter is
compensation scheme. At start-up the induction motor speed is 0 and vector control
219
Chapter 7
algorithm introduces a slip to start acceleration and the converter output electrical
Figure 7-20 Converter capacitor voltages (top) and arm current (bottom) at start of acceleration
At higher frequencies the converter is supplying power to the electric motor and in the
constant power region operates with a fixed boost ratio. A zoomed capture of the
220
Experimental validation
The high effective switching frequency is difficult to capture, but the line-to-line
voltage steps can be clearly seen. The converter operates with a high boost ratio, with the
peak voltage being 4 times higher than the dc source. At the same time the dc-side current
has consistent ripple amplitude. The cause for the small oscillations in the dc-side current
are unknown at this point, but are suspected to be caused by the OPAL-RT digital
controller measurement.
During braking, the BMCI must invert the motor iq current and starts recovering part
of the kinetic energy of the flywheel. An oscilloscope capture is shown in Figure 7-22:
221
Chapter 7
the converter. Before and immediately after the start of braking the converter is operating
with a voltage boost, but a lower factor than during acceleration. The dc-side current
When the converter frequency goes below 5Hz the low-frequency balancing control
system is enabled and the 100Hz balancing current is injected, as shown in Figure 7-23.
The arm current before the transition has very small dc value, as the losses in the electric
motor do not allow for much energy to be regenerated. In addition, the 5Hz point was
chosen as practically at this frequency the peak capacitor voltage ripple was equal to the
222
Experimental validation
Figure 7-23 Transition from normal operation to low-frequency ripple compensation mode
The capacitor voltage ripple is directly affected by the motor rated power and base
frequency. In Figure 7-24 another traction cycle is shown, together with the voltage of a
223
Chapter 7
The converter is operating in low-frequency mode up to time = 1.146s, where the stator
frequency is below 8 Hz. After the initial response to the motor torque step at t=0s, the
ripple in that region is less than the peak value of the cycle. Maximum ripple occurs at
t=3s, where the BMCI enters field weakening mode. After this point the amplitude of the
capacitor instantaneous power stays constant, while frequency increases, thus reducing
the peak-to-peak voltage ripple. The maximum value measured is 10.5V. Using the
displayed in the surface plot in Figure 7-25. Therefore, the experimental results confirm
the validity of the theoretical analyses and the practicality of the proposed converter as a
traction drive.
224
Experimental validation
Figure 7-25 Surface plot of capacitor voltage ripple as a function of power factor and output frequency
7.4. SUMMARY
This chapter has been focussed on the experimental validation of the proposed
converter. The hardware that has been designed for the project has been presented as a
test rig with both passive and active loads. The experiments show a very good agreement
with the analytical model and the numerical simulations and demonstrate that the BMCI
is attractive for traction drives of electric vehicles, especially those with a low voltage dc
power supply.
Experiments with a variable passive load show the basic operations of the converter
with and without closed-loop control, exploring in detail the buck-boost characteristics of
The converter has been then connected to an induction motor and measurements have
been presented for acceleration and braking cycles for all 4 quadrants of operation. These
experiments show that the converter can operate in the full speed range required by a
traction motor, while operating at maximum torque and providing voltage boost ratios up
225
Chapter 7
to 4.5. The converter shows superior voltage waveforms, compared to those of a 2-level
converter, even though it uses only 2 sub-modules per arm, while showing lower
capacitor voltage ripple at low frequencies, compared to a HB MMCC. At full power, the
converter has very low circulating current ripple, even though it uses a relatively small
While low-frequency operations introduce a higher dc-side current ripple, this effect
can be mitigated by using some form of energy storage either at the converter’s dc-side
or integrated in each sub-module. This would likely be the case of hydrogen fuel cell
vehicles, as energy storage would be determinant to reduce the peak power and, hence,
the cost of the fuel cell stack. The proposed converter would be then particularly useful
for this application, as the presence of low voltage sub-modules with floating capacitors
is particularly suited for the connection of low voltage dc energy storage like
226
Conclusion and future possibilities
This thesis has been focussed on the analysis of the Boost Modular Cascaded Inverter
for railway traction systems. The converter topology is essentially identical to a basic full-
bridge modular cascaded converter. However, this PhD project has explored a completely
new way of using this circuit – as a traction inverter with buck and boost voltage
operations.
The principle of operation of the converter has been described in details and the
mathematical model has been derived analytically. This work has presented the detailed
dc transfer function of the BMCI that has been validated using open-loop experiments,
showing the natural response of the converter. As conventional MCCs are used as buck
converters or they are used to interface networks with fixed voltages, the maximum boost
It has also been shown that the BMCI can be modelled as a combination of a cascaded
boost converter and a 3-phase inverter, which facilitates the design of the main control
systems. This project has also developed a new balancing algorithm for the sub-module
A converter design methodology has also been proposed to calculate the required
converter inductance on the basis of the modulation scheme, the number of sub-modules,
and the mean capacitor voltage. This has been used to design a 160 kW BMCI
asynchronous motor drive that can be used for a fuel-cell powered tram. This setup is
227
Chapter 8
Finally, the thesis has presented the experimental results from a laboratory prototype
of the converter. The small-scale BMCI has been setup to drive an asynchronous machine
with a peak line-to-line voltage of 200 V, while drawing power from a 50 V dc power
supply. To simulate a traction load, the electric motor’s shaft has been mechanically
coupled to a solid steel flywheel. This setup simulates an inertial load and provides
roughly 6-7 seconds of acceleration time at 500 W machine power. The logged data has
been analysed and the following conclusions and recommendations can be made.
8.1. CONCLUSIONS
8.1.1. Suitability of BMCI as a boost traction inverter
This thesis has shown that the experimental prototype can operate in all 4 quadrants,
i.e. it can both supply and recover power from the mechanical load. During the
acceleration, the converter maintains constant current amplitude, while increasing line-
to-line voltage. Above the asynchronous machine base frequency, the converter adjusts
the field current to achieve constant power output achieving the required field-
weakening. When a braking command is applied, the BMCI reverses the power flow,
starting from the field-weakening region and moving to the constant torque region below
the base frequency. The most challenging operation point for the BMCI occurs when the
rotor frequency is close to 0, where ωe →0, and capacitor voltage ripple increases
significantly. The experimental results show that the BMCI can compensate for the voltage
ripple with appropriate control of the zero-sequence voltage and the peak-to-peak value
As with any other type of boost converter, the BMCI parasitic resistances need to be
taken into account in order to define the correct steady-state gain and to avoid unstable
228
Conclusion and future possibilities
boost operations. The analytical model shows that large parasitic resistance can severely
limit the maximum boost ratio. The model has been validated using practical experiments,
using a known passive resistive-inductive load. As expected, the converter voltage gain is
substantially smaller than the ideal value, as the prototype converter uses overrated
The research has found that the maximum required capacitor voltage is not defined by
the minimum dc-side voltage, but by the maximum. The experimental converter operates
with a boost ratio of roughly 0.14 during motoring, and 0.18 during regenerative braking.
As the sub-module capacitor voltages are regulated to a constant value, the modulation
index has more headroom during motoring, where dc-side voltage is lower. This
counterintuitive result makes the BMCI suitable for applications like dc-powered rail
vehicles, where intermittent line voltages are often encountered. In such circumstances,
the high converter capacitance can be utilised to maintain constant output power.
Chapter 5 has presented the methodology for the design of the converter. It is shown
that the total converter capacitance is dependent only on the boost ratio and the machine
transistor switching frequency, the power factor, and the motor current amplitude.
High speed motor drives are an area where multilevel inverters are preferred, as they
typical 2-level inverters operate with switching frequencies as low as 500Hz. The BMCI is
229
Chapter 8
The BMCI produces a higher quality voltage waveform, compared to an MCC with the
same number of sub-modules, n, but lower than a flying capacitor or neutral point clamp
multilevel inverter. However, the BMCI also removes the need for a boost converter, and
square of the number of sub-modules. Thus, the BMCI can be a power dense solution for
certain applications and it can also achieve higher efficiency and better motor
performance.
sufficient depth. Some of the more interesting ones are indicated in the following.
Permanent magnet synchronous motors offer slightly higher efficiency and power
level inverter collapses the voltage induced by the magnet can be harmful for the power
electronics. The BMCI is capable of operating with 0 dc-side voltage, as long as there is a
power source connected to the 3-phase side like the case of permanent magnet motor
drives and, hence, there would not be issues in case of a dc-link fault;
A recent paper has shown that this current component can be injected to reduce
capacitor voltage ripple, resulting in smaller capacitance requirement for the converter.
The proposed study will have to investigate the maximum 2nd harmonic current
amplitude that does not violate the peak arm current requirement. This value will
0 has explored the effects of the phase shift angle between the carriers of arm a and
that of arm b, θcarrier. It has been shown that, depending on number of sub-modules, the
converter can achieve either m or 2m-1 output levels. However, the value of θcarrier that
achieves the lowest inverter ac-side voltage THD causes the highest current ripple in the
circulating current. Further work is required to fully describe the performance of popular
This thesis has described the effects of conventional modulation methods on the
design of the leg inductor. Depending on the chosen scheme, the frequency spectrum of
the circulating dc current shows different power spread. As conventional MCCs are buck
converters, they use ddc,x = 0.5, resulting in a boundary condition for the modulation
waveforms. The effect of dc duty cycle that is not divisible by n have not yet been explored.
Moreover, there are other modulation methods, where the PWM carriers have variable
phase, dependent on modulation depth. Future work can explore optimum PWM carrier
Traction converters must achieve both high power density and high efficiency. Typical
power inductors for traction are large, heavy and have a high cost. The proposed BMCI
can reduce the required inductance by increasing the number of sub-modules, but the
total converter capacitance will be higher than a 2-level inverter with the same power
rating. A detailed model of BMCI power density must be constructed in order to estimate
a realistic power density figure for a traction converter using the proposed topology.
231
Chapter 8
Efficiency of the converter has not been investigated. With a conventional boost
converter, high boost ratios result in high converter losses, as the transistors switch high
currents at a high voltage. In a BMCI, the number of sub-modules is arbitrary, and the sub-
module voltage can be chosen. For example, low-voltage MOSFETs can achieve very low
an n-module converter, with 2n devices, can have a lower on resistance than a transistor
A comprehensive analysis of converter can show how efficiency scales with different
number of sub-modules, taking into account the type of device, required blocking voltage
The effects of dc duty cycle ddc,x on ac voltage THD must be explored further. Lower
ddc,x increases voltage boost and the number of effective inverter ac-side voltage levels,
and results in lower THD. At the same time, however, one cross-pair of transistors in a
sub-module bridge will have much longer conduction times and any distortion, caused by
very high current application, diode rectifiers are a very popular solution due to the low
power dissipation, low voltage drop and high power factor. However, diode rectifiers have
a regulation effect due to the dc-side inductance and they require very comprehensive
output short circuit protection, as diodes are not controllable devices. Thyristors, on the
232
Conclusion and future possibilities
other hand, offer reverse blocking capabilities, but introduce high number of harmonics
into the 3-phase power supply and reactive power when the voltage output is regulated.
A BMCI converter can be used as an active rectifier, that offers both high power factor
and fault management capability. A power electronics converter can implement arbitrary
fault responses and the system current limit can be set to a much lower value than that
for diode rectifiers. As an additional functionality, a BMCI traction rectifier will allow for
bidirectional power flow, where braking energy from a train slowing down can be
The BMCI is a modular structure, consisting of many identical modules. Several papers
have shown that MCC converters can integrate energy storage devices like batteries and
supercapacitors, separately controlling energy stored, and power drawn from the dc-link.
This principle can be applied to a BMCI hydrogen-battery hybrid drive that does not
require a battery management system and can fully utilise the energy stored in the battery
banks.
Examining this application will require incorporating a battery model into the model
described in chapter 3. The control system of the BMCI must be modified to manage the
energy stored in the batteries, and the sub-module voltage balancing must be converted
The design of a hybrid BMCI drive would be a very interesting topic. While it has been
shown that an MCC can have a sub-module for every battery, such a converter with a
standard voltage of 300V will require hundredths of sub-modules. Moreover, the device
utilisation will be poor, as a typical lithium-ion battery cell has a terminal voltage that
varies between 3.5V and 4.2V. At the same time commercial low-voltage/high-current
MOSFET brake-down voltages are 30V or higher. As an example, to fully utilise a 30V
233
Chapter 8
transistor, the BMCI sub-module will require 4 batteries, producing up to 16.8V. This
allows sufficient margin for a fast high-current semiconductor. The hybrid BMCI must be
designed according to a predetermined traction cycle, in order to set the maximum stored
energy requirement. The same amount of energy can be stored in either low number of
battery cells with high capacity, or a high number of low-capacity cells. From the results
of this project, it would be possible to determine the optimum number of battery cells,
and sub-modules that can achieve balance between converter efficiency, complexity, and
power density.
The BMCI must also control the dc-side current to be always positive, which is a
requirement of hydrogen fuel cells. A braking chopper could be used to limit the peak
battery power during regenerative braking but is expected that it will not be required.
234
BMCI Design
BIBLIOGRAPHY
[1] J. Young, "150 YEARS OF THE LONDON UNDERGROUND," Tunnels & Tunnelling
International, p. 3.
[2] R. Dennis, "Making the Underground Underground," The London Journal, vol. 38,
pp. 203-225, 2013/11/01 2013.
[3] P. Dawson, "Electric Railways and Tramways, their Construction and Operation: A
Practical Handbook," in Cambridge Library Collection - Technology, ed Cambridge:
Cambridge University Press, 2013, pp. 1-17.
[4] K. O. Anderson, "Wanted: A modern diesel-electric rail car," Transactions of the
American Institute of Electrical Engineers, Part II: Applications and Industry, vol. 72,
pp. 345-349, 1953.
[5] S. E. Steinbrook, "Diesel locomotives for lightweight passenger trains," Electrical
Engineering, vol. 76, pp. 483-483, 1957.
[6] L. Greer, "Modern Electrical Equipment for Industrial Diesel-Electric
Locomotives," Transactions of the American Institute of Electrical Engineers, vol. 61,
pp. 229-232, 1942.
[7] G. Huldschiner, "Railway electrification, diesel-electric locomotives, and some
future aspects of electric traction," Transactions of the American Institute of
Electrical Engineers, Part II: Applications and Industry, vol. 71, pp. 48-50, 1952.
[8] F. A. Harper, "Diesel-electric locomotives of the British Railways," Journal of the
Institution of Electrical Engineers, vol. 1950, pp. 319-320, 1950.
[9] T. J. Woods, "The preference for diesel-electric locomotives," Electrical
Engineering, vol. 68, pp. 1030-1030, 1949.
[10] G. B. Anderson and A. J. Peters, "An overview of the maintenance and reliability of
AC traction systems," in Proceedings of the 1993 IEEE/ASME Joint Railroad
Conference, 1993, pp. 7-15.
[11] S. Guo, Z. Rong, J. Yao, and H. Wang, "Reliability modeling and assigning method for
HXD electric locomotive," in 2013 International Conference on Quality, Reliability,
Risk, Maintenance, and Safety Engineering (QR2MSE), 2013, pp. 289-295.
[12] C. R. Avery, "Power electronics reliability in rail traction," in IEE Colloquium on
Power Electronics Reliability - Promise and Practice. Does it Deliver? (Digest No.
1998/202), 1998, pp. 6/1-6/7.
[13] C. Duchesne, T. Lebey, M. Mermet-Guyennet, E. Dutarde, and S. Dagdag,
"Contribution to the stress grading in integrated power modules," in 2007
European Conference on Power Electronics and Applications, 2007, pp. 1-9.
[14] M. Andresen, K. Ma, G. Buticchi, J. Falck, F. Blaabjerg, and M. Liserre, "Junction
Temperature Control for More Reliable Power Electronics," IEEE Transactions on
Power Electronics, vol. 33, pp. 765-776, 2018.
[15] A. R. Miller, K. S. Hess, D. L. Barnes, and T. L. Erickson, "System design of a large fuel
cell hybrid locomotive," Journal of Power Sources, vol. 173, pp. 935-942,
2007/11/15/ 2007.
[16] H. Hoffmann, G. Jacobs, A. Kramer, and C. Pelger, "Evaluation of the efficiency
improvement potential of diesel multiple unit railcars by the use of an electrical
power split hybrid drive," in Drive Systems 2017 7th VDE/VDI Symposium, 2017,
pp. 1-6.
235
Appendix A
[17] A. Jaafar, C. R. Akli, B. Sareni, X. Roboam, and A. Jeunesse, "Sizing and Energy
Management of a Hybrid Locomotive Based on Flywheel and Accumulators," IEEE
Transactions on Vehicular Technology, vol. 58, pp. 3947-3958, 2009.
[18] R. Cousineau, "Development of a hybrid switcher locomotive the Railpower Green
Goat," IEEE Instrumentation & Measurement Magazine, vol. 9, pp. 25-29, 2006.
[19] N. Shiraki, H. Satou, and S. Arai, "A hybrid system for diesel railcar series Ki-Ha
E200," in The 2010 International Power Electronics Conference - ECCE ASIA -, 2010,
pp. 2853-2858.
[20] U. Sarma and S. Ganguly, "Determination of Rating Requirement for Fuel-Cell-
Battery Hybrid Energy System to Substitute the Diesel Locomotives of Indian
Railway," in 2017 14th IEEE India Council International Conference (INDICON),
2017, pp. 1-6.
[21] A. Lajunen, "Evaluation of Diesel and Fuel Cell Plug-In Hybrid City Buses," in 2015
IEEE Vehicle Power and Propulsion Conference (VPPC), 2015, pp. 1-6.
[22] A. Hoffrichter, A. R. Miller, S. Hillmansen, and C. Roberts, "Well-to-wheel analysis
for electric, diesel and hydrogen traction for railways," Transportation Research
Part D: Transport and Environment, vol. 17, pp. 28-34, 2012/01/01/ 2012.
[23] M. Tekin, D. Hissel, M. Pera, and J. M. Kauffmann, "Energy-Management Strategy for
Embedded Fuel-Cell Systems Using Fuzzy Logic," IEEE Transactions on Industrial
Electronics, vol. 54, pp. 595-603, 2007.
[24] Y. M. A. Welaya, M. M. El Gohary, and N. R. Ammar, "A comparison between fuel
cells and other alternatives for marine electric power generation," International
Journal of Naval Architecture and Ocean Engineering, vol. 3, pp. 141-149,
2011/06/01/ 2011.
[25] S. Dehn, M. Woehr, and A. Heinzel, "Development of a near-dead-ended fuel cell
stack operation in an automotive drive system," in 2011 IEEE Vehicle Power and
Propulsion Conference, 2011, pp. 1-6.
[26] J. P. Gagge, "PEM fuel cells versus diesel generators — which solution to pick?," in
INTELEC 2008 - 2008 IEEE 30th International Telecommunications Energy
Conference, 2008, pp. 1-5.
[27] B. Martin, J. Gerfried, W. Ewald, B. Walter, K. Manfred, S. Mario, et al., "Hydrogen
powered fuel cell forklifts — Demonstration of green warehouse logistics," in 2013
World Electric Vehicle Symposium and Exhibition (EVS27), 2013, pp. 1-4.
[28] M. Bakowski, Lim, J-K., Kostov, K., "Green PE O2.1 Transnational Technology and
Product Roadmap," Green Power Electronics, Product Roadmap25/01/2018
2018.
[29] J. W. Kolar, Marz, M., Wolfgang, E., Round, S. D., "Automotive Power Electronics
Roadmap," in International Conference of Integraded Power Electronics Systsems
(CIPS), Nuremberg, Germany, 2008.
[30] E. ERTRAC, ETIP SNET, "European Roadmap Electrification of Road Transport,"
European Road Transport Research Advisory Council, European Road Transport
Research Advisory Council Website2017.
[31] EETT, "Electrical and Electronics Technical Team Roadmap (2017)," U.S. DRIVE
Partnership, U.S. Office of ENERGY EFFICIENCY & RENEWABLE ENERGY2017.
[32] K. Olejniczak, T. McNutt, D. Simco, A. Wijenayake, T. Flint, B. Passmore, et al., "A
200 kVA electric vehicle traction drive inverter having enhanced performance
over its entire operating region," in 2017 IEEE 5th Workshop on Wide Bandgap
Power Devices and Applications (WiPDA), 2017, pp. 335-341.
[33] G. Sarriegui, S. Beushausen, and R. W. D. Doncker, "Comparison of high frequency
inductors for bidirectional DC-DC converters for electric vehicles," in 2016 18th
236
BMCI Design
237
Appendix A
238
BMCI Design
[65] A. Weschta, "Traction technology of the new class S252 and S447 power units for
the Spanish Railways Renfe," in PESC '92 Record. 23rd Annual IEEE Power
Electronics Specialists Conference, 1992, pp. 411-416 vol.1.
[66] S. Kimura, Y. Itoh, W. Martinez, M. Yamamoto, and J. Imaoka, "Downsizing Effects
of Integrated Magnetic Components in High Power Density DC-DC Converters for
EV and HEV Applications," IEEE Transactions on Industry Applications, vol. 52, pp.
3294-3305, 2016.
[67] L. Arioua and M. Ameziani, "Coupled-Inductors Losses Modelling for Size and
Weight Optimization Process Avoiding Time-Consuming Co-Simulations," in PCIM
Europe 2018; International Exhibition and Conference for Power Electronics,
Intelligent Motion, Renewable Energy and Energy Management, 2018, pp. 1-6.
[68] W. Martinez, S. Kimura, J. Imaoka, M. Yamamoto, and C. A. Cortes, "Volume
comparison of DC-DC converters for electric vehicles," in 2015 IEEE Workshop on
Power Electronics and Power Quality Applications (PEPQA), 2015, pp. 1-6.
[69] J. W. Kolar, U. Drofenik, J. Biela, M. L. Heldwein, H. Ertl, T. Friedli, et al., "PWM
Converter Power Density Barriers," in 2007 Power Conversion Conference - Nagoya,
2007, pp. P-9-P-29.
[70] W. Martinez, C. Cortes, M. Yamamoto, J. Imaoka, and K. Umetani, "Total volume
evaluation of high-power density non-isolated DC–DC converters with
integrated magnetics for electric vehicles," IET Power Electronics, vol. 10, pp. 2010-
2020, 2017.
[71] D. Rahman, A. J. Morgan, Y. Xu, R. Gao, W. Yu, D. C. Hopkins, et al., "Design
methodology for a planarized high power density EV/HEV traction drive using SiC
power modules," in 2016 IEEE Energy Conversion Congress and Exposition (ECCE),
2016, pp. 1-7.
[72] A. B. Ponniran, K. Orikawa, and J. Itoh, "Minimum Flying Capacitor for N-Level
Capacitor DC-DC Boost Converter," IEEE Transactions on Industry Applications, vol.
52, pp. 3255-3266, 2016.
[73] O. Hegazy, J. V. Mierlo, and P. Lataire, "Analysis, Modeling, and Implementation of
a Multidevice Interleaved DC/DC Converter for Fuel Cell Hybrid Electric Vehicles,"
IEEE Transactions on Power Electronics, vol. 27, pp. 4445-4458, 2012.
[74] B. Vulturescu, A. D. Bernardinis, R. Lallemand, and G. Coquery, "Traction power
converter for PEM fuel cell multi-stack generator used in urban transportation," in
Power Electronics and Applications, 2007 European Conference on, 2007, pp. 1-10.
[75] M. Hirakawa, M. Nagano, Y. Watanabe, K. Andoh, S. Nakatomi, and S. Hashino, "High
power density DC/DC converter using the close-coupled inductors," in 2009 IEEE
Energy Conversion Congress and Exposition, 2009, pp. 1760-1767.
[76] S. E. Schulz, "Exploring the High-Power Inverter: Reviewing critical design
elements for electric vehicle applications," IEEE Electrification Magazine, vol. 5, pp.
28-35, 2017.
[77] S. Narasimhan, M. Karami, R. Tallam, and M. Das, "Evaluation of SiC based inverter
drives," in 2017 IEEE 5th Workshop on Wide Bandgap Power Devices and
Applications (WiPDA), 2017, pp. 1-8.
[78] T. Yoneyama, T. Yamamoto, K. Kondo, T. Furuya, and K. Ogawa, "Fuel cell powered
railway vehicle and experimental test results," in 2007 European Conference on
Power Electronics and Applications, 2007, pp. 1-10.
[79] S. Salehahari, R. Pashaei, E. Babaei, and C. Cecati, "Coupled-inductor based
multilevel inverter," in 2017 14th International Conference on Electrical
Engineering/Electronics, Computer, Telecommunications and Information
Technology (ECTI-CON), 2017, pp. 907-910.
239
Appendix A
240
BMCI Design
[94] P. Palanivel and S. S. Dash, "Analysis of THD and output voltage performance for
cascaded multilevel inverter using carrier pulse width modulation techniques,"
IET Power Electronics, vol. 4, pp. 951-958, 2011.
[95] L. G. Franquelo, J. Rodriguez, J. I. Leon, S. Kouro, R. Portillo, and M. A. M. Prats, "The
age of multilevel converters arrives," IEEE Industrial Electronics Magazine, vol. 2,
pp. 28-39, 2008.
[96] A. Antonopoulos, L. Angquist, S. Norrga, K. Ilves, L. Harnefors, and H. P. Nee,
"Modular Multilevel Converter AC Motor Drives With Constant Torque From Zero
to Nominal Speed," IEEE Transactions on Industry Applications, vol. 50, pp. 1982-
1993, 2014.
[97] J. J. Jung, H. J. Lee, and S. K. Sul, "Control Strategy for Improved Dynamic
Performance of Variable-Speed Drives With Modular Multilevel Converter," IEEE
Journal of Emerging and Selected Topics in Power Electronics, vol. 3, pp. 371-380,
2015.
[98] A. Edpuganti, A. K. Rathore, and B. M. Joshi, "Optimal low-switching frequency
pulsewidth modulation of dual modular multilevel converter for medium-voltage
open-end stator winding induction motor drive," in IECON 2016 - 42nd Annual
Conference of the IEEE Industrial Electronics Society, 2016, pp. 3751-3755.
[99] D. C. Ludois and G. Venkataramanan, "Modular multilevel converter as a low
inductance machine drive," in 2012 IEEE Power and Energy Conference at Illinois,
2012, pp. 1-4.
[100] M. Winkelnkemper, A. Korn, and P. Steimer, "A modular direct converter for
transformerless rail interties," in 2010 IEEE International Symposium on Industrial
Electronics, 2010, pp. 562-567.
[101] J. Kolb, F. Kammerer, M. Gommeringer, and M. Braun, "Cascaded Control System of
the Modular Multilevel Converter for Feeding Variable-Speed Drives," IEEE
Transactions on Power Electronics, vol. 30, pp. 349-357, 2015.
[102] J. Kucka, D. Karwatzki, L. Baruschka, and A. Mertens, "Modular Multilevel
Converter with Magnetically Coupled Inductors," IEEE Transactions on Power
Electronics, vol. PP, pp. 1-1, 2016.
[103] J. Qin, M. Saeedifard, A. Rockhill, and R. Zhou, "Hybrid Design of Modular Multilevel
Converters for HVDC Systems Based on Various Submodule Circuits," IEEE
Transactions on Power Delivery, vol. 30, pp. 385-394, 2015.
[104] B. Li, R. Yang, D. Xu, G. Wang, W. Wang, and D. Xu, "Analysis of the Phase-Shifted
Carrier Modulation for Modular Multilevel Converters," IEEE Transactions on
Power Electronics, vol. 30, pp. 297-310, 2015.
[105] B. P. McGrath, C. A. Teixeira, and D. G. Holmes, "Optimized Phase Disposition (PD)
Modulation of a Modular Multilevel Converter," IEEE Transactions on Industry
Applications, vol. 53, pp. 4624-4633, 2017.
[106] K. Haridas, S. Khandelwal, and A. Das, "Three phase to single phase modular
multilevel converter using full bridge cells," in 2016 IEEE International Conference
on Power Electronics, Drives and Energy Systems (PEDES), 2016, pp. 1-5.
[107] M. Vasiladiotis, N. Cherix, and A. Rufer, "Single-to-three-phase direct AC/AC
modular multilevel converters with integrated split battery energy storage for
railway interties," in Power Electronics and Applications (EPE'15 ECCE-Europe),
2015 17th European Conference on, 2015, pp. 1-7.
[108] R. Zeng, L. Xu, L. Yao, and B. W. Williams, "Design and Operation of a Hybrid
Modular Multilevel Converter," IEEE Transactions on Power Electronics, vol. 30, pp.
1137-1146, 2015.
241
Appendix A
242
BMCI Design
244
BMCI Design
[154] Y. Tang, L. Ran, O. Alatise, and P. Mawby, "Capacitor Selection for Modular
Multilevel Converter," IEEE Transactions on Industry Applications, vol. 52, pp.
3279-3293, 2016.
[155] R. Ferrero, M. Marracci, and B. Tellini, "Single PEM Fuel Cell Analysis for the
Evaluation of Current Ripple Effects," IEEE Transactions on Instrumentation and
Measurement, vol. 62, pp. 1058-1064, 2013.
[156] C. y. A. d. F. CAF, S.A. (2018, Jan 2018). URBOS tram range. Available:
http://www.caf.net/en/productos-
servicios/familia/urbos/descripcion/gama_urbos.php
[157] H. S. Hansen, M. U. Nawaz, and N. Olsson, "Using operational data to estimate the
running resistance of trains. Estimation of the resistance in a set of Norwegian
tunnels," Journal of Rail Transport Planning & Management, vol. 7, pp. 62-76,
2017/06/01/ 2017.
[158] ABB, "Product note: Modular induction traction motors," ABB, Ed., ed. ABB traction
motors page: ABB, 2011.
[159] C. Restrepo, Konjedic, T., Garces, A., Calvente, J., Giral, R., "Identification of a Proton-
Exchange Membrane Fuel Cell's Model Parameters by Means of an Evolution
Strategy," IEEE Transactions on Industrial Informatics, vol. 11, pp. 548-559, 2015.
[160] A. M. Dhirde, N. V. Dale, H. Salehfar, M. D. Mann, and T. H. Han, "Equivalent Electric
Circuit Modeling and Performance Analysis of a PEM Fuel Cell Stack Using
Impedance Spectroscopy," IEEE Transactions on Energy Conversion, vol. 25, pp.
778-786, 2010.
[161] H. Corporation, "HYPM-HD90 Power Module," ed, 2016.
[162] J. Kim, J. Lee, and B. H. Cho, "Equivalent Circuit Modeling of PEM Fuel Cell
Degradation Combined With a LFRC," IEEE Transactions on Industrial Electronics,
vol. 60, pp. 5086-5094, 2013.
[163] J. Hua, Xu, L., X. Lin, and M. Ouyang, "Modeling and experimental study of PEM fuel
cell transient response for automotive applications," Tsinghua Science and
Technology, vol. 14, pp. 639-645, 2009.
[164] KY, "Technical information IGBT-Module FF300R07ME4_B11," 3.1 ed.
infineon.com: Infineon Technologies, 2014.
[165] N. Nithil Babu, "Comparison of Weighted Total Harmonic Distortion of Various
PWM Techniques employed in Multi-level Inverters," International Journal of
Advanced Research in Electrical, Electronics, and Instrumentation Engineering, vol.
4, pp. 9246-9251, November 2015 2015.
[166] D. Energia, "Power Electronics Capacitors," ed, 2010, pp. 19,20.
[167] SBE, "SBE Technology for Traction Drive Inverters," SBE Inc, SBE website2014.
[168] T. v. d. S. Steeg, J. (2016, Aluminium electrolytic capacitors for traction
applications. 2. Available:
https://www.vishay.com/docs/28412/aluelectrfortracapps.pdf
[169] R. Burkart and J. W. Kolar, "Component cost models for multi-objective
optimizations of switched-mode power converters," in 2013 IEEE Energy
Conversion Congress and Exposition, 2013, pp. 2139-2146.
[170] C. W. S. Inc. (2013, 23.08). Very High Current Power Inductor 450 Amperes Rating
Shoebox Series. Available:
http://www.coilws.com/index.php?main_page=index&cPath=208_212
[171] L. C. Magnetics. (2016, 23.08). GAPPED INDUCTOR. Available:
https://www.lcmagnetics.com/inductors/gapped-inductor/
245
Appendix A
246
BMCI Design
This appendix details the design of a BMCI for an urban tram. The converter
capacitance is designed according to the machine parameters from Table 6-2 and fuel-cell
The induction machine line-to-line voltage is 532VRMS, which results in required peak
ac-side voltage of at least 750V line-to-line. From equation (5-8) the required arm voltage
is:
of sub-modules, with derating the converter can use 600 or 650V transistors and a total
of n=2 sub-modules per arm. These are industry standard voltages and there are multiple
possible choices for transistors, both IGBTs and MOSFETs. The peak transistor current is:
VIN (A-3)
d dc , x =
= 0.132
2 2VLL
The arm dc current can be inferred from the fuel cell output current rated power:
Iin (A-4)
I dc, x = = 366.7 A
3
The coincidence duty cycle can be calculated from (5-16), for ddc,x<0.5/n:
Setting a design limit of <20% current ripple, equal to roughly 75A per leg, and with
interleaved arm a and b carriers, using equation (5-17) the required inductance becomes:
247
Appendix A
1
Lleg = dcoinc (1 − dcoinc )VCTsw,eff 249 H (A-6)
idc , x
The leg inductor is centre-tapped to reduce the converter output inductance, and the
sufficiently large to maintain the designed LM at current equal to the peak dc current. The
248
Control system design
The control system of the BMCI is designed according to the parameters found in Table
6-4. The converter operating point was found in Chapter 7 to be ddc,x=0.165 and Dac,x = 0.8.
( )
rSM = ddc , x 2 + dac , x 2 rC + 2rSW = 63.1m (B-7)
The phase resistance Rx can be derived from the motor rated voltages and currents, as
Rx
2 + rL + nrSM (B-9)
cos
Rx , SM = = 3.52
nDac , x , RMS 2
The bode plot of the transfer function, including zero-order hold delay, of the
circulating current idc,x with input dc duty cycle ddc,x, equation (4-26), is shown below:
249
Appendix B
From equations (4-27) through (4-29) the zero frequency, natural frequency, and dc
2 (B-10)
z = = 56.73rad / s
RSM C
2nVC
Gidc, x ( 0 ) = − = 3.17 103 A / V (B-12)
nd RSM + rarm
2
loop bandwidth ω0dB,CL=2π400 rad/s. The current compensator gains kp,i and ii,i are
250
Control system design
idcGidcn 2 (B-13)
i ,0 dB = = 4.09 106 rad / s
z
i ,CL (B-14)
k p ,idc = = −0.0015
i ,0 dB
The bode plot of the open loop transfer function can be seen in :
The stability of the system is checked by evaluating the open loop Nyquist response:
251
Appendix B
The Nyquist plot shows that the system is stable, with phase margin of 65° and gain
margin of 13.8dB. The plot of the closed-loop transfer function is shown below:
The bode response of capacitor voltage transfer function from equation (4-41) can now
be plotted:
252
Control system design
= = 0.1980V / A (B-16)
Gidc 2nDdc , x
The dc-gain of the capacitor transfer function is less than 0dB, but equation (4-44)
computes the asymptotic intersection of the break frequency and the 0dB line. Thus, the
Gvc
v ,0dB = z = 114.17rad / s (B-17)
Gidc
The closed-loop crossover frequency of the voltage loop is set one decade below that
of the current controller, and the compensator gains , using (4-45) and (4-46), are:
v ,0 dB ,CL (B-18)
k p ,vc = = 55.03
v ,0 dB
z (B-19)
ki ,vc = k p ,vc = 317.32
2
The stability of the system can be evaluated using a Nyquist plot:
253
Appendix B
The phase margin was found to be 64 degrees and the gain margin of 6.5dB.
254
Converter PCB schematics
255
Appendix C
256
Converter PCB schematics
257
Appendix C
258