Você está na página 1de 127

UNIVERSITY OF CALGARY

Improving Inverter Efficiency at Low Power Using a Reduced Switching Frequency

by

Tahsina Hossain Loba

A THESIS
SUBMITTED TO THE FACULTY OF GRADUATE STUDIES
IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE
DEGREE OF MASTER OF SCIENCE

GRADUATE PROGRAM IN MECHANICAL AND MANUFACTURING


ENGINEERING

CALGARY, ALBERTA

SEPTEMBER, 2015

Tahsina Hossain Loba 2015


i

Abstract

The inverter is a major component of a renewable energy system and its performance affects
the overall performance of the system. For typical household applications in rural areas, often
there is need to operate at low power conditions where inverter efficiency can drop
dramatically. Efficient operation at low power is important especially for stand-alone solar
systems in developing countries where system cost must be kept low. In this thesis, the impact
of switching frequency upon switching loss for a SPWM inverter is investigated. Results,
from mathematical modeling, simulation and experimental implementation, show the same
trend that reducing the switching frequency reduces switching loss at low power levels thus
improves inverter efficiency. This may result in a reduced PV module size requirement and
thus lower system cost. The inverter proposed in the thesis operates efficiently at low power
(e.g. 9W) as well as at rated power conditions (e.g. 200W).

ii

Acknowledgements

In the name of Allah, the Most Beneficent, the Most Merciful.

I am highly grateful to my supervisor, Dr. David Wood, and co-supervisor Dr. Ed Nowicki
for their kind support, invaluable advice, readiness to have meetings frequently throughout
the period I have been doing research work. I am thankful to them for believing in me and
supporting me morally and intellectually when I was struggling with my experiments.

I am also thankful to my colleagues for their continuous inspiration.

I am most thankful to Allah who helped me at every stage during the course of my research
work and life.

iii

Dedication

I dedicate this thesis work to my beloved husband, Wali, who constantly supported me
through the entire time I spent in this masters program. Without his help and support, it
would not have been possible to complete the degree.

iv

Table of Contents

Abstract ............................................................................................................................... ii
Acknowledgements ............................................................................................................ iii
Dedication .......................................................................................................................... iv
Table of Contents .................................................................................................................v
List of Tables .................................................................................................................... vii
List of Figures and Illustrations ....................................................................................... viii
List of Symbols, Abbreviations and Nomenclature .............................................................x
CHAPTER ONE: INTRODUCTION ..................................................................................1
1.1 Problem Statement .....................................................................................................1
1.2 Background and Motivation ......................................................................................2
1.3 Objective of the Thesis ..............................................................................................4
1.4 Scope of the Thesis ..................................................................................................5
1.5 Thesis Outline ..........................................................................................................6
CHAPTER TWO: LITERATURE REVIEW ......................................................................8
2.1 DC/AC Power Inverter ............................................................................................8
2.2 Categorizing Power Inverters ..................................................................................8
2.2.1 Modified Sine Wave Inverters ...........................................................................9
2.3 Full Bridge Inverter ...............................................................................................11
2.4 Pulse Width Modulation ........................................................................................12
2.4.1 SPWM Switching Techniques .........................................................................15
2.4.1.1
SPWM with Bipolar Switching ..............................................................15
2.4.1.2
SPWM with Unipolar Switching ............................................................17
2.5 Advantages of using Unipolar SPWM...................................................................20
2.6 Overview of the Related Work ..............................................................................20
2.7 Chapter Summary ..................................................................................................24
CHAPTER THREE: INVERTER LOSS MODEL AND DESIGN CONSIDERATIONS25
3.1 Introduction ............................................................................................................25
3.2 Full Voltage Design and Scaled Down Design .....................................................26
3.3 Proposed Inverter Topology ..................................................................................27
3.4 Inverter Loss Components based on Loss Model ..................................................28
3.4.1 Conduction Losses ...........................................................................................30
3.4.2 Switching Losses .............................................................................................34
3.4.3 IGBT Gate Drive Losses..................................................................................36
3.5 Inverter Efficiency Calculation ..............................................................................37
3.6 Chapter Summary ..................................................................................................43
CHAPTER FOUR: SIMULATION RESULTS FOR PROPOSED FULL BRIDGE
INVERTER ...............................................................................................................44
4.1 Schematic of the Full Bridge Inverter System under Study ....................................44
v

4.2 Simulation Circuit Description ................................................................................45


4.2.1 SPWM Controller ............................................................................................45
4.2.2 Gate Drive Circuit ............................................................................................47
4.2.3 IGBT Switching Circuit ...................................................................................48
4.2 Different Load Conditions .......................................................................................51
4.3 Simulation Results ...................................................................................................56
4.4 Chapter Summary ....................................................................................................60
CHAPTER FIVE: EXPERIMENTAL RESULTS ...........................................................61
5.1 Introduction..............................................................................................................61
5.2 Hardware Overview ...............................................................................................62
5.2.1 Power Supply for Transistor Drive Circuit ......................................................65
5.2.2 Isolation Circuit ...............................................................................................65
5.2.3 Optocoupler Circuit Operation ........................................................................66
5.2.4 Full Bridge Inverter Circuit .............................................................................68
5.2.5 Snubber Circuit ................................................................................................70
5.2.6 RC Snubber ......................................................................................................70
5.3 Load Modeling and Load Value Calculations .......................................................73
5.3.1 Determination of Component values for Light Load (CFL) Model ................74
5.3.2 Heavy Load Model Calculations .....................................................................78
5.3.3 Current Sense Resistors ..................................................................................82
5.4 Experimental Results .............................................................................................82
5.4.1 Measurement Approach ...................................................................................83
5.4.2 Results and Discussion ....................................................................................84
5.5 Chapter Summary ..................................................................................................86
CHAPTER SIX: CONCLUSIONS ...................................................................................88
6.1 Summary ................................................................................................................88
6.2 Contributions .........................................................................................................89
6.3 Suggestions for Future Work .................................................................................91
REFERENCES ..................................................................................................................93
APPENDIX A: SCHEMATIC DIAGRAM OF FULL BRIDGE INVERTER ...............102
APPENDIX B: COMPONENT LIST FOR EXPERIMENTAL SETUP ........................104
APPENDIX C: MICROCONTROLLER (MSP430G2553) OVERVIEW .....................107
APPENDIX D: MICROCONTROLLER CODE ............................................................112

vi

List of Tables
Table 3.1 Power Loss Calculation under Different Load Conditions for Variable
Switching Frequency (for 388.96Vdc, Rated Power 200W) .................................... 38
Table 3.2 Efficiency under Different Load Conditions for Variable Switching
Frequency (for 388.96Vdc, Rated Power 200W)...................................................... 39
Table 3.3 Power Loss Calculation under Different Load Conditions for Variable
Switching Frequency (for 25Vdc, Rated Power 12.856W) ...................................... 41
Table 3.4 Efficiency under Different Load Conditions for Variable Switching
Frequency (for 25Vdc, Rated Power 12.856W)........................................................ 42
Table 4.1 Unipolar SPWM Switching Logic .................................................................... 50
Table 4.2 Simulation Results Under Different Load Conditions For Variable Switching
Frequency (For 388.75Vdc) ...................................................................................... 57
Table 4.3 Simulation Results Under Different Load Conditions For Variable Switching
Frequency (For 25Vdc) ............................................................................................. 58
Table 5.1 Hardware Results of Full load and Light load .................................................. 84
Table 5.2 Comparison of the efficiencies ......................................................................... 86

vii

List of Figures and Illustrations


Figure 2.1 Modified Sine Wave, Sine Wave, Square Wave Inverter Output Waveforms . 9
Figure 2.2 Modified Sine Waveform ................................................................................ 10
Figure 2.3 Full Bridge Inverter Topology ........................................................................ 11
Figure 2.4 (a) Comparison of Triangular Wave and Sinusoidal Wave (total time period
= 20ms, 1ms per division (horizontally), 0.1V per devision (vertically) ) ............... 14
Figure 2.4 (b) Resultant SPWM Wave (total time period = 20ms, 1ms per division
(horizontally), 0.1V per devision (vertically) ) ......................................................... 14
Figure 2.5 Comparator Used to Generate the Bipolar SPWM Signals ............................. 16
Figure 2.6 Waveform for SPWM with Bipolar Voltage Switching (a) Comparison of
Reference and Triangular Waveform (b) Output Waveform with Sinusoidal
Fundamental Component Shown by Dashed Line.................................................... 16
Figure 2.7 Two Comparators Generating Unipolar SPWM Signal .................................. 17
Figure 2.8 Waveform for SPWM with Unipolar Voltage Switching (a) Comparison of
Reference and Triangular Waveform (b) Gating Pulses for A + and B- (c) Gating
Pulses for A- and B + (d) Output Waveform ............................................................ 18
Figure 3.1 Block Diagram of Inverter System Including Controller ................................ 28
Figure 3.2 IGBT Vce and Rq (where, Rq = Vce/iq) at Rated Load Condition; note
that points plotted here are taken from the Device Datasheet .................................. 31
Figure 3.3 Diode Vd and Rd (where, Rd = Vak/id) at Rated Load Condition; note
that points plotted here are taken from the Device Datasheet .................................. 31
Figure 3.2 IGBT Vce and Rq (where, Rq = Vce/iq) at Light Load Condition; note
that points plotted here are taken from the Device Datasheet .................................. 32
Figure 3.3 Diode Vd and Rd (where, Rd = Vak/id) at Rated Load Condition; note
that points plotted here are taken from the Device Datasheet .................................. 32
Figure 4.1 Schematic of Controller for generating SPWM gating signals as used in
PSpice Simulations ................................................................................................... 46
Figure 4.2 Simulated Waveforms for SPWM with Unipolar voltage switching (a)
Sinusoidal Reference waveform and Triangular Carrier waveform, (b), (c),(d),(e)
gating pulses for A +, A-, B +and B- respectively after Sine and Triangular
comparison ................................................................................................................ 47
viii

Figure 4.3 PSpice Modelling of IGBT Gate Drive Circuit. (a) PSpice Gain Circuit
which is also called an E Block. (b) Op Amp Symbol. (c) Op Amp model. ............ 48
Figure 4.4 Simulated Output Voltage Waveform of the inverter at 2.5kHz for SPWM
with the Unipolar Switching ..................................................................................... 49
Figure 4.5 Schematic of Equivalent Circuit of CFL Model connected as a load at the
inverter output ........................................................................................................... 52
Figure 4.6 Simulated Current Waveform from the Equivalent Circuit of CFL Model
connected as a Load at the Inverter output (Total Time Interval = 30ms, 1ms time
division (horizontally), 10mA current division (Vertically)) ................................... 54
Figure 4.7 Simulated Voltage Waveform from the Equivalent Circuit of CFL Model
connected as a load at the inverter output (Total Time Interval = 20ms, 1ms time
division (horizontally), 50V voltage division (Vertically)) ...................................... 54
Figure 4.8 Schematic of Equivalent Circuit of CFL Model connected as a load at the
inverter output ........................................................................................................... 55
Figure 4.9 Simulated Current Waveform from the Equivalent Circuit of Heavy Load
Model connected as a load at the inverter output (Total Time Interval = 30ms, 1ms
time division (horizontally), 1A current division (Vertically)) ................................ 55
Figure 4.10 Simulated Voltage Waveform from the Equivalent Circuit of Heavy Load
Model connected as a load at the inverter output (Total Time Interval = 20ms, 1ms
time division (horizontally), 50V voltage division (Vertically)) .............................. 56
Figure 5.1 Partial Schematic of the Experimental Inverter System .................................. 64
Figure 5.2 Internal Circuit of FOD3184 IC ...................................................................... 67
Figure 5.3 Block Diagram showing Optocoupler Interface .............................................. 68
Figure 5.4 Equivalent Circuit of the Snubber Circuit ....................................................... 71
Figure 5.5 Equivalent Circuit Model of CFL for Scaled Down Version .......................... 73
Figure 5.6 Equivalent Heavy Load for Scaled Down Version ......................................... 78
Figure A-1 Full Bridge Inverter Supplying Power to the Heavy Load. ......................... 102
Figure B-1 DC-DC Converter NME0515SC .................................................................. 104
Figure B-2 DC-DC Converter NME0515SC .................................................................. 105
Figure B-3 Optocoupler FOD3184 ................................................................................. 106
ix

List of Symbols, Abbreviations and Nomenclature


Symbol

Definition

()

()
()

()
()

()

Bottom left switch


Upper left switch
Bottom right switch
Upper right switch
DC load capacitor
Snubber capacitor
Load capacitor
Total gate capacitance
Duty cycle
Total energy dissipated
Turn off energy loss IGBT
Turn on energy loss IGBT
Total energy loss IGBT
Total energy dissipated
Frequency of control signal
Frequency of carrier signal
Switching frequency
Closed circuit current
Collector Current
LED forward current
Worst case peak current
Total current through load circuit
RMS current flowing through diode
DC current for charging circuit
Peak load current
RMS current flowing through IGBT
RMS output current
Test current IGBT
Output current in time domain
Input current in time domain
Correction factor
Amplitude modulation index
Frequency modulation index
Emitter power loss
Total gate driver loss
Internal circuitry power loss
Input power
Output power
Total conduction loss IGBT
x

()
()

Total conduction loss diode


Total switching loss
Total gate charge
Load resistor
Sense resistance
Diode on state resistance
Collector to emitter on state resistance
Apparent power
Turn on voltage of IGBT
Diode forward voltage
Collector to emitter voltage
Diode forward voltage
Collector to emitter voltage
IGBT on state voltage
Diode on state forward voltage
Open circuit voltage
Bus voltage
LED forward voltage
IGBT on state voltage
Positive DC Voltage
Negative DC Voltage
Supply voltage
Emitter Voltage
Gate to emitter voltage
Triangular carrier signal
Reference sinusoidal signal
Input AC system
Rectified voltage
RMS output voltage
Total supply voltage
Test voltage IGBT
Output voltage in time domain
Input voltage in time domain
Peak magnitude of control signal
Peak magnitude of carrier signal
Voltage difference between collector and emitter
Voltage difference of diode forward voltage
Capacitive reactance
Total impedance
Total CFL phase angle
Total heavy load phase angle

xi

Chapter One: Introduction

Much of the worlds population is without electricity. Solar energy is a sustainable means of
providing electricity, and is especially suitable for locations where it is difficult or too expensive
to construct high voltage transmission lines. The efficiency of a solar energy system largely
depends on the efficiency of the system power inverter. This thesis addresses the inverter
efficiency issue at low power levels for rural stand-alone solar home systems. Chapter one
describes the research problem, motivation behind the research, contributions and brief outline of
the next chapters.

1.1 Problem Statement

In the case of a stand-alone rural solar home system in a developing country, the typical load
profile can vary between 0W to 200W [1]. Under optimal conditions for a rural stand-alone solar
home system, the inverter may operate at around 85%-95% efficiency while operating at its
maximum rated power. 200W is considered as the maximum rated power in the context of this
thesis. But the efficiency of the same inverter drops significantly while operating under 20% of its
maximum rated load [2]. This is a challenging problem for a stand-alone solar home system in a
developing country where the household load profile remains at low power for most of the time
during the day and at night [3]. The inverter efficiency at low power is investigated and improved
in this thesis which may improve the overall usefulness of the solar home system. Specifically,
9W (4.5% of maximum rated power 200W) is considered as an example of low power that might
1

be used to operate a Compact Fluorescent Lamp (CFL). The improvement in inverter efficiency
may result in decreased cost of the overall solar home system. Previous research [3], [24], [25],
[26], [27], [28], [29], [30] does not sufficiently address the low power (i.e. 5W-10W) inverter
efficiency problem found in rural solar home systems in developing countries. To address the
problem in this thesis, a 200W inverter is simulated and a scaled down 12.856W version is
implemented experimentally. The 12.856W version was chosen for quick implementation, but
operated at RMS current level of the full voltage design. The proposed inverter has acceptable
efficiency even at low power levels with a simple control circuit and without the need of a filter
circuit at the inverter output. However, it is yet to be determined which loads can be operated
without the output filter.

1.2 Background and Motivation

Due to geographical and socio-political challenges, over 1.2 billion people or 20% of the world's
population, are without access to electrical grid power, almost all of whom live in the developing
countries. Hence, the development of micro-grids and stand-alone power systems is seen as an
economic way to raise the living standard for remote villages which are not connected to the grid.
One of the renewable sources that receives much attention is solar photovoltaic (PV) power. Every
hour the sun radiates more energy onto the earths surface than is consumed globally in one year
[4]. To harness the power of solar energy, improvements in the efficiency of photovoltaics and
electrical storage are required to reduce the variability and intermittency of solar power. So,
considering a scenario where there is no electric grid connection and ample sunlight, a stand-alone
solar home system is often preferable to other forms of renewable energy. The main components
2

of a typical solar home system are solar module(s), charge controller, battery, DC-DC converter
and the inverter. One of the essential parts of the stand-alone solar home system is a single-phase
full-bridge inverter. An inverter is necessary because the majority of the electrical appliances run
on 220 (applicable for developing countries in general). The inverter is needed to convert the
DC output of the solar module or battery into usable AC power.

In a typical developing world context, residential electrical demand is highly dynamic and stays at
low power levels for about 20 hours in a day [3]. For a rural solar home system, the load demand
remains at low power (1W to 10W) for a significant amount of time, especially at night. Improving
inverter efficiency at low power means more efficient use of the renewable energy resource.
Conversion efficiency is a prime consideration for all switch-mode power supplies (SMPSs). It is
even more critical for solar home systems, where prolonging battery life is a key goal, especially
at night and for cloudy days, when there is little or no solar energy supply from the PV modules.
In addition, if the energy input to the inverter can be efficiently converted at low power, it is
suggested in this thesis that it may also be possible to reduce the PV module area. In both cases
this may result in savings of the system capital cost.

At low power, switching loss becomes a significant portion of total power that is wasted. Thus,
the switching loss is mainly responsible for the decrease in efficiency of the inverter during low
power operations. One way to reduce switching loss at low power is to reduce the switching
frequency which also results in decreased power quality. The term power quality is used in
reference to the voltage, current and frequency of the power delivered [5]. For the purposes of this
thesis, high power quality means a sinusoidal inverter output voltage of nominal RMS voltage at
3

the nominal power frequency. Power quality is an issue in grid connected inverters and there are
stringent requirements to comply with standards (e.g. IEEE Std 1547 is typical of such standards).
Also, the IEEE [6] and the IEC [7] standards put limitations on the maximum allowable amount
of injected dc current into the grid [8]. However, the power quality of a stand-alone solar home
system is not subject to these constraints. In this thesis, it is suggested that a reduction in switching
frequency at low power levels, and some degradation in power quality, may result in a significant
improvement in inverter efficiency at low power levels. Another motivating factor is that higher
switching loss not only results in reduced energy efficiency but also exerts more stringent
requirements on the thermal management for the switching devices.

1.3 Objective of the Thesis

The objective of the thesis is to develop an inverter for low power applications (on the order of
50W to 500W rated power) with high efficiency operation over a wide range of power levels (i.e.
down to about 5% of the rated power). One example of a low power load is a Compact Florescent
Lamp (CFL) which may consume only 5W to 10W depending on the light output desired. As noted
above, in this thesis the inverter is rated for a power output of 200W (this is also referred to as the
maximum rated power) and low power operation of that inverter is considered to be 9W or 4.5%
of 200W. Therefore, the objective is to design an inverter that operates efficiently both at low
power and maximum rated power. Using simulation, a 200W inverter prototype is designed which
has about 85% efficiency even when operated at 9W utilizing inexpensive and reliable components
available in a developing country.

1.4 Scope of the Thesis

In this research, inverter efficiency is improved for low power operation by reducing switching
frequency. Inverter switching loss (the dominant loss component at low power) is a function of
switching frequency and the current flowing through the switching device [3]. The choice of the
switching frequency involves a trade-off between requirements for high efficiency, sinusoidal
output power quality if a filter is employed, and low cost [8]. At rated power (200W in this context)
there are several kinds of losses that play a significant role in terms of decreasing the inverter
efficiency [5]. Low power means less current, less current means less loss but switching losses
increase in proportion to the switching frequency for a given constant load level [9], [10]. That
means, reduction of switching loss at low power is the key to improved inverter efficiency.
Therefore, to address the research problem, switching frequency is reduced to improve the inverter
efficiency at low power. In addition, inverter output characteristics and power quality are also
considered while operating at a reduced switching frequency. The reduced switching frequency
does reduce power quality, but it is expected that most household loads will tolerate this reduced
power quality. The scope of the research includes the following:

1. Inverter loss model and efficiency calculation based on mathematical analysis.


2. Inverter simulation in OrCad PSpice to observe efficiency for heavy and light load profiles.

3. Experimental implementation of the scaled down version (of rated power 12.856W) for
quick prototype development to observe inverter efficiency.

4. Analysis of efficiency trends by comparing efficiencies found in mathematical analysis,


simulation and experimental implementation.

1.5 Thesis Outline

The remainder of the thesis is organized as follows:

Chapter 2 briefly presents the relevant literature and describes recent research to improve inverter
efficiency at low power. A classification of different kinds of inverters commonly used in a standalone solar home system is presented. The detailed working principles of the modified sine wave
inverter and the reason for using this inverter is explained in detail. The full-bridge inverter
topology is also discussed in this chapter. The switching pattern of the full- bridge inverter using
the USPWM (Unipolar Sinusoidal Pulse Width Modulation) technique is presented in a table to
explain the states of each transistor while switching. The chapter also describes and critiques
previous research in the area of low power inverter efficiency improvement.
Chapter 3 details a mathematical analysis of the inverter loss components. Transistor and diode
model parameters are calculated using datasheet information. The process of obtaining the
parameters is a modification of a technique found in the literature. Based on the equations
presented in this chapter, a table has been provided showing the individual loss components and
the efficiencies for different switching frequencies. This chapter also presents the design
considerations and topology of the proposed inverter. The reason behind using a hard switched
inverter instead of using a soft switched is explained in detail.
6

Chapters 4 presents the computer simulation results of the proposed inverter. The simulation
results are discussed and analyzed in terms of efficiency for different switching frequency and load
levels. Equivalent load models of the heavy load and light load conditions are implemented in
PSpice to observe the output voltage and current characteristics of the inverter. A similar table to
that in chapter 3, showing the total loss and efficiency measurements is presented in this chapter.

Chapter 5 describes experimental implementation of the scaled down version of the proposed
inverter. It begins with an overview of the overall inverter prototype and how different blocks
interact with each other to provide AC at the inverter output. The description of the selection of
the overall inverter circuit components is presented. A detailed calculation of the value of the
components used to build the equivalent heavy (12.856W) and light (0.578W) load is also provided
in this chapter. A comparative analysis of the three results (mathematical, simulation,
experimental) is also presented in a table to show the agreement among them. Based on the
operation of the scaled down inverter version (i.e. 12.856W), it is suggested that it is worthwhile
to proceed with the full voltage design (i.e. 200W) in the near future.

Chapter 6 presents conclusions based on the mathematical model, simulation model and
experimental results. Also, future work directions are suggested.

Chapter Two: Literature Review

2.1 DC/AC Power Inverter

Power inverters are devices which convert DC to AC. The purpose of a DC/AC power inverter is
to take the DC power supplied by a battery or solar module and transform it into the standard AC
output (220 at 50Hz in the context of developing countries in general, e.g. Bangladesh
residential power). The solar modules and the batteries can store and supply only DC power but
most household appliances and other electrical equipments require AC input power to perform. To
supply the AC power to the household appliances the inverter is an essential part of the solar home
system [8].

2.2 Categorizing Power Inverters

There are three different types of power inverter output waveforms; square wave, modified sine
wave and pure sine wave as shown in Figure 2.1. These inverter output waveforms differ,
providing varying levels of distortion that can affect electronic devices in different ways. Modified
sine wave inverters have a lower Total Harmonic Distortion (THD) than a square wave inverter
but higher THD than a sine wave inverter [11]. THD measures how much the power waveform is
distorted by the harmonics. For running typical household appliances a modified sine wave
inverter is a reliable and cost-effective choice. Though the modified sine wave inverter does not
produce a true AC sine wave power, it does provide an affordable option and for many power

applications is perfectly adequate. Modified sine wave inverters approximate a sine wave and have
low enough harmonics that they do not cause problems with typical household equipment [11].

Modified Sine wave sits at zero for a


certain time then rises or falls

Modified Sine Wave


Sine wave
Square Wave

Figure 2.1 Modified Sine Wave, Sine Wave, Square Wave Inverter Output Waveforms

The modified sine wave inverter costs half the price of the sine wave inverter [11]. In the context
of a rural stand-alone solar home system, where power quality is not a regulated requirement,
modified sine wave inverters provide affordable and portable AC power [12].

2.2.1 Modified Sine Wave Inverters

In the modified sine wave inverter, there are three voltage levels in the output waveform, positive,
negative, and zero as shown in Figure 2.2, with a dead zone between the positive and negative
pulses [11]. Modified sine wave inverters can be designed to satisfy the efficiency requirements
of the photovoltaic system while being less expensive than pure sine waveform inverters [12].
These inverters are capable of operating a wide variety of loads; electronic and household items
including, CFLs, filament bulbs, TV, VCR, satellite receiver, computer, refrigerator, sewing

machine etc. [13]. Thus, most of the household appliances commonly used in a developing country
will work satisfactorily with a modified sine wave inverter [11].

Figure 2.2 Modified Sine Waveform

Since modified sine wave inverters have higher THD than pure sine wave inverters, they are not
suitable for applications where the power quality requirement is stringent (e.g. grid connected solar
systems). Grid connected power inverters must comply with the appropriate standards for THD
where modified sine wave inverters do not [6], [7]. After choosing the modified sine wave inverter
the next step is to determine suitable topologies and the control techniques for the proposed
inverter.

10

2.3 Full Bridge Inverter

Figure 2.3 displays the basic inverter circuit using the full bridge topology. A full bridge inverter
is a switching configuration composed of four switching devices (e.g. IGBT switches) in an
arrangement that resembles an H shape (hence the alternative name H-bridge) [13]. The Hbridge circuit may be designed for a particular modified sine wave inverter application. By
controlling the switches in the bridge by controller signals, a positive, negative, or zero potential
voltage can be placed across the bridge output i.e. between A and B, as shown in Figure 2.3. The
switching and control techniques are described in detail in section 2.3.

B+

A+

VDC

+
_

Load

B
B-

A-

Figure 2.3 Full Bridge Inverter Topology

The switches used to implement a full bridge configuration can be mechanical or built from solid
state transistors, though mechanical switches are almost no longer in use. Normally, Bipolar
Junction Transistors (BJT), Metal Oxide Semiconductor Field Effect Transistors (MOSFET) or
Insulated Gate Bipolar Transistors (IGBT) devices are used as switches, each type having its own
advantages and disadvantages. In this thesis, the IGBT has been used as a switching device due to
11

its high-current handling capability and suitability for varying load, low duty cycle (ON time of a
pulse divided by the total switching period) and low frequency applications. IGBTs can operate
within a wide range of switching frequencies and are very easily configurable for varying
switching frequency applications. Generally, the IGBT has superior conduction characteristics
compared to the MOSFET. In this thesis, a low loss IGBT is chosen as the switching device as it
can operate over a wide range of switching frequencies (<40kHz in hard switching) [14], [15].

2.4 Pulse Width Modulation

There are various ways to control the switches of a full bridge inverter and generate the desired
inverter output. In power electronic converters, Pulse Width Modulation (PWM) is extensively
used as a means of powering the AC devices from a DC source [16]. Variation of the duty cycle
(if the switch is pulsed ON at variable durations, the duty cycle varies) [18] in the PWM signal is
used to provide a specific voltage pattern that will appear to the load as an AC signal [17]. The
pattern at which the duty cycle of a PWM signal varies, can be created through simple analog
components, a digital microcontroller, or specific PWM integrated circuits [13]. Most commonly,
PWM signals are generated through the microcontroller due to the ease of implementation, long
term-reliability and precise switch timing [14]. Some other techniques are also used for high
frequency switching (e.g. multilevel switching, Pulse Frequency Modulation, Pulse Amplitude
Modulation, etc.) but due to its many advantages, a PWM technique is usually chosen over other
switching techniques.

12

PWM is the process of modifying the width of the pulses in a pulse train in direct proportion to a
control signal (for example, the greater the control voltage, the wider the resulting pulses) [15].
When the control signal is sinusoidal the process is called Sinusoidal Pulse Width Modulation
(SPWM). By using a sinusoid at the desired frequency as the control voltage for a SPWM circuit,
it is possible to produce a high power waveform whose average voltage varies in a sinusoidal
manner suitable for driving the semiconductor devices used for switching. The analog SPWM
control approach requires the generation of both the reference and the carrier signals that feed into
a comparator which creates output signals based on the difference between the signals [14]. As
mentioned earlier, the reference signal is sinusoidal with the frequency of the desired output signal,
while the carrier signal is often either a saw tooth or triangular wave at a frequency significantly
greater than the reference frequency. The frequency of the reference signal determines the inverter
power output frequency and the reference peak amplitude controls the modulation index and the
RMS value of the output voltage [16]. This process is shown in Figure 2.4(a) with the triangular
carrier wave in green, the sinusoidal reference wave in red and the resultant SPWM pulses in blue
in 2.4(b).

13

Figure 2.4 (a) Comparison of Triangular Wave and Sinusoidal Wave (Total Time Period =
20ms, 1ms per Division (Horizontally), 0.1V per Devision (Vertically) )

Figure 2.4 (b) Resultant SPWM Wave (Total Time Period = 20ms, 1ms per Division
(Horizontally), 0.1V per Devision (Vertically) )

14

2.4.1 SPWM Switching Techniques

The two types of SPWM switching techniques are unipolar and bipolar switching to create either
a unipolar or bipolar output at the load. The control signals depend on comparing a reference signal
and carrier signal [19], [20]. They are now discussed in turn.

2.4.1.1 SPWM with Bipolar Switching

This technique uses a comparator as shown in Figure 2.5, to compare the reference voltage
waveform with the triangular carrier signal as shown in Figure 2.6 (a). In SPWM
with bipolar switching, the H-bridge output voltage swings between + and as shown in
Figure 2.6(b).

The switching scheme that will implement bipolar switching using the full bridge inverter, as
shown in Figure 2.6, is determined by comparing the instantaneous reference and carrier signals
[19]:

+ and are on when > ( = +)

(2.3)

+ and are on when < ( = )

(2.4)

15

VControl
VCarrier

+
A+ and B-

A- and B+
Not

Figure 2.5 Comparator Used to Generate the Bipolar SPWM Signals

Figure 2.6 Waveform for SPWM with Bipolar Voltage Switching (a) Comparison of
Reference and Triangular Waveform (b) Output Waveform with Sinusoidal Fundamental
Component Shown by Dashed Line
16

2.4.1.2 SPWM with Unipolar Switching

In this scheme, the triangular carrier waveform ( ) is compared with two control signals
( ) which are positive and negative signals. The basic circuit to produce SPWM with
unipolar voltage switching is shown in Figure 2.7. The difference between the bipolar SPWM and
unipolar SPWM generators is that the latter uses another comparator to compare between the
inverse reference waveform shown in Figure 2.7. The comparison of these two signals
produces the unipolar voltage switching signal. The output waveform is switched either from high
(+) to zero or from low () to zero as shown in Figure 2.8 (a). The gating pulses of the
four IGBT switches and output waveform are shown in Figure 2.8 (b), (c) and (d). The effective
switching frequency seen by the load is doubled that of gating signal. Due to this, the harmonic
content of the output voltage waveform is reduced compared to bipolar switching. In Unipolar
switching, the amplitude of the significant harmonics and its sidebands is much lower for all
modulation indexes [19].

VControl

+
A+

BNot
VCarrier

+
VControl

A-

B+

Not

Figure 2.7 Two Comparators Generating Unipolar SPWM Signal

17

The switching scheme to implement unipolar switching using the full bridge inverter as shown in
Figure 2.8 is determined by comparing the instantaneous reference and carrier signals [20], [21]:

+ is on when >

(2.5)

is on when <

(2.6)

+ is on when >

(2.7)

is on when

(2.8)

<

Figure 2.8 Waveform for SPWM with Unipolar Voltage Switching (a) Comparison of
Reference and Triangular Waveform (b) Gating Pulses for + and (c) Gating Pulses for
and + (d) Output Waveform

The number of pulses per half-cycle depends upon the ratio of the frequency of the carrier signal
( ) to the modulating sinusoidal signal ( ). The frequency of the control signal (i.e.
of the modulating signal) sets the inverter output frequency ( ) and the peak magnitude of
control signal controls the modulation index (ratio of the peak magnitude of control signal,
18

to the peak magnitude of carrier signal, ) which in turn controls the RMS output
voltage. If is the width of pulse, the RMS output voltage can be determined by [19]:

= (2
=1

2 1/2
)

(2.9)

The amplitude modulation index is defined as:

(2.10)

where, = peak magnitude of control signal (modulating sine wave)


= peak magnitude of the carrier signal (triangular signal)

The frequency modulation ratio is defined as:

(2.11)

where, = frequency of control signal (sinusoidal wave)


= frequency of carrier signal (triangular wave)

19

2.5 Advantages of using Unipolar SPWM

The unipolar SPWM voltage switching scheme is selected in this thesis because this method offers
the advantage of effectively doubling the switching frequency of the inverter voltage. A particular
advantage of the unipolar SPWM approach is that, this method reduces the harmonics in the single
phase inverter [22], [23]. That means, selecting unipolar SPWM as the switching scheme in the
proposed inverter is appropriate as there is no filter at the inverter output.

2.6 Overview of the Related Work

Researchers have tried to improve inverter efficiency at low power in many different ways. Five
different and typical techniques have been identified from the large number of techniques in the
literature. These techniques are:

1. Using hybrid switches (combination of MOSFET and IGBT) [3].

2. Enabling pulse skipping mode (operating the inverter at the maximum efficiency point for
shorter intervals) [26].

3. Implementing a combination of hybrid pulse width modulation (HPWM) and zero voltage
switching (ZVS) to improve the efficiency of the inverter for high switching frequency
applications [27].
20

4. Resonant mode switching methods can be used such as the LLC burst mode for gridconnected applications (where there is a slight improvement in efficiency for light load
related to a small change in switching frequency) [28], [29], [30].

5. Using variable switching frequency in discontinuous current mode (DCM) (variable


switching frequency control method allows extending the input voltage range
considerably) [31].

In [3], parallel IGBT-MOSFET switch operation is analyzed and it has been shown that light load
efficiency can be improved with hybrid switch use. The faster response of the MOSFET to
switching signal commands, compared to the IGBT, is used for minimizing IGBT turn-off losses.
This paper suggests that if such a hybrid switch is employed it is better to use the MOSFET only
for light loads. The lowest power level is 100W and the efficiency is around 78%. However,
consideration must be given to protect the light load switch (MOSFET/IGBT) from a sudden
increase in load current.

In [26] a pulse skipping control strategy is developed to improve inverter efficiency at low power.
The pulse skipping technique is preferable in grid tied inverters. When the input power drops below
a certain level, the inverter can be controlled to stop feeding power into the grid continuously.
Thus, enter the pulse-skipping operation mode. This strategy is not suitable for stand-alone inverter

21

systems due to its complex control circuitry and with this control method the efficiency drops
significantly while operating under 30W (where the inverter is rated for 200W).

[27] implements hybrid pulse width modulation (HPWM), and zero voltage switching (ZVS)
together, to improve the efficiency of the inverter for high switching frequency applications. The
HPWM scheme while operating with ZVS during positive half cycle of the output frequency
reduces the switching loss to approximately one half of the loss with a standard bipolar PWM
technique. However, the efficiency reduces significantly if IGBT switches are chosen instead of

the MOSFET switches in the HPWM inverter. Because, the ZVS technique while applied in a
HPWM inverter, cannot reduce the losses due to the IGBT tailing current losses. Also, the
additional circuitry (containing: parasitic capacitor, resonant inductor, DC-link capacitor, etc.)
with an additional DC-link switch in order to implement the ZVS technique for ZVS operation
adds some power loss [27]. Efficient operation can only be achieved for a very small load range.
The HPWM prototype only works at the high end of the switching frequency range (50kHz to
180kHz) and suffers from high switching losses.

In [28] burst mode control along with synchronous rectifier (SR) is applied to improve light load
efficiency of an LLC resonant converter. The LLC resonant converter refers to a unique
combination of two inductors and one capacitor (L-L-C) in the integrated transformer stage
before the inverter stage [25], [29]. In this technique, LLC series resonant converter is employed
in order to achieve zero voltage switching (ZVS) low turn on and turn off current in the full bridge
inverter configuration on primary side of the transformer. While in the half bridge configuration
of the secondary side, zero current switching (ZCS) is achieved to maintain low turn on and turn
22

off current. Thus, suffers from large switching loss, conduction loss and core loss at light load
because of the magnetizing current which flows reversely from secondary to primary. To address
this large losses at light load condition, later a burst mode control was applied in order to block
the switching driver signals periodically and thus limiting the power conversion only during the
time having switching driver signals. Thus, the driver loss and the switching loss were reduced at
light load. This method is only applicable for improving inverter efficiency at 10% of the rated
load condition, and the ripple voltage and burst mode losses increase significantly at around 5%
of the rated load. The LLC series resonant converter needs to be further modified while operating
at rated load conditions. In order to address, high voltage ripple at the inverter output at around
10% of the rated load and to reduce losses during burst mode operation, a Capacitor-InductorCapacitor (CLC) output filter is applied. This technique fails to perform efficiently if the
capacitor values are not chosen to satisfy the ripple requirements at different load points. As
mentioned above, this technique also requires a resonant transformer stage before the inverter
stage, which will add to the overall loss of the system along with increased cost [29]. The switching
frequency range is maintained in a relatively narrow range and is applicable only when the
switches are controlled through a soft switching method. Further, the overall topic and analysis is
still considered complex, and is poorly understood [30].
In [31] a variable switching frequency control method is proposed to address the low input voltage
range and high conduction loss issue in a discontinuous conduction mode (DCM) fly-back
micro-inverter. This technique is again more preferable for grid connected inverter system, focuses
on reducing the conduction loss in order to improve the overall efficiency and the output current
is reduced significantly. Also, in this technique switching frequency is in the range of 40kHz to
100kHz and operates within a range of 260W to 80W.
23

2.7 Chapter Summary

Based on the literature review, it can be concluded that the techniques investigated so far for
improving inverter efficiency at low power are mostly done at power levels higher than those of
interest here, have complex and expensive control circuitry, and are applicable for high switching
frequency applications. As the proposed inverter is not connected to the grid some techniques are
not suitable. At low power, switching losses dominate all other losses in the inverter system and it
has been demonstrated that switching loss is a function of inverter switching frequency [32].
Therefore, the technique proposed in this thesis to improve inverter efficiency at low power by
reducing switching frequency is a simple and appropriate solution for a stand-alone solar home
system with light load household appliances.

24

Chapter Three: Inverter Loss Model and Design Considerations

3.1 Introduction

To reduce switching loss at low power levels for a single phase SPWM stand-alone inverter, the
switching frequency is reduced [78]. Decreasing the switching frequency to improve the overall
inverter efficiency is possibly the simplest of all the techniques. The objective of this thesis is to
design a 200W inverter which is capable of performing at about 85% efficiency even when
operated at 9W (4.5% of its rated power). The proposed design does not include a filter, to reduce
the cost and to keep the design simple. The critical design aspect considered in this context is to
allow distortion of the load current without hampering operation (such as a CFL light or cell phone
charger) at lower switching frequency.

During mathematical modelling, the switching frequency is varied from 20kHz to 200Hz to
observe the change in inverter efficiency. The maximum efficiency at low power is found by
dividing the output power by input power at 200Hz which is the minimum switching frequency in
this context. While operating at low switching frequency the load has to deal with higher total
harmonic distortion of the inverter output current and voltage waveforms which results in low
power quality at the inverter output. Hence, it is suggested here that a reduction in switching
frequency at low power levels and some degradation in power quality, may result in a significant
improvement in inverter efficiency at low power levels.

25

3.2 Full Voltage Design and Scaled Down Design

In this thesis two inverter designs are considered. The full voltage design is rated for 200W power
output. The intended application is in a rural home in the developing world. Many countries have
a 220V RMS AC standard at 50Hz. This is the standard used in this thesis. For a 200W power
output, a 220V RMS sinusoidal voltage has a peak value of 2x220V=311V. For the 200W design,
an 80% modulation index is chosen (in case the DC supply drops, the modulation index can be
increased to compensate for the drop). Thus the nominal dc supply voltage is 311V/0.80 = 389V.
The next step is to find the RMS output current when the inverter is operating in the heavy load
and light load scenarios. The formula for calculating the RMS output current of the inverter is,
_ = / ( ). Thus, the RMS current output of the inverter output at
heavy load is, 200W/(220V*0.8) = 1.136A, where, the output power is 200W, the RMS output
voltage is 220V and the power factor is 0.8 (corresponding to an inductive household load, e.g.
table fan, small refrigerator, sewing machine or combination of these). Similarly, the RMS current
output of the inverter output at light load is, 9W/(220V*0.65) = 62.9mA, where, the output power
is 9W, the RMS output voltage is 220V and the power factor is 0.65. The reason for choosing a
smaller power factor for light load scenario (e.g. 9W CFL) is explained in section 5.3.1 of
Chapter 5.

The experimental implementation of the inverter is based on a scaled down version of the 200W
design (for quicker implementation). This scaled down design is based on a dc supply voltage of
25V (available from a lab power supply). Thus, for a modulation index of 80%, the peak of the
sinusoidal output voltage is 25Vx0.80 = 20V, giving a RMS output voltage of 20V/2 = 14.142V.
26

The output current rating for the scaled down design is kept equal to that of the full-voltage design.
For the heavy load scenario where the rated power is 12.856W (i.e. 200W x scaling factor, where
the scaling factor is 20V/2V / 220V = 0.0642824) and the RMS output voltage is 14.142V, given
a modulation index of 0.8, the RMS current rating is 12.856W/(14.142V*0.8) = 1.136A (as
expected, since voltage is scaled but current is not). For the light load scenario where the output
power is 0.57825W (i.e. 4.5% of 12.8566W) and the RMS output voltage is 14.142V, given a
modulation index of 0.65, the RMS current rating is 0.57825W/(14.142V*0.65) = 0.0629A or,
62.9mA (again, as expected). The detailed explanation of the reasons behind using a lower power
factor (0.65) for light load (CFL) is explained in section 5.3.1 of Chapter 5. As mentioned before,
although the RMS voltage has been scaled down from 220V to 14.142V, the RMS current values
(i.e. 1.136A for heavy load and 62.9mA for light load) and power factors (i.e. 08 for heavy load
and 0.65 for light load) are maintained to be the same for the full voltage design and the scaled
down version.

3.3 Proposed Inverter Topology

As shown in Figure 3.1 a hard switched unipolar SPWM single phase full bridge inverter was
simulated and a prototype of the proposed inverter was built. A hard switched inverter is
appropriate in this context because of its inexpensive implementation, simple control circuitry and
ease of maintenance [33]. Unipolar SPWM has been chosen as the modulation technique because
the inverter does not have any filter and unipolar SPWM can better approximate a sinusoid
compared to the case of bipolar SPWM [20]. The full bridge topology is chosen with
considerations that it must be capable of delivering high current at low voltage. This high current
27

at low voltage property is important if the inverter is designed for photovoltaic applications [34].
The standard for output current THD (<100%) is maintained for both rated and light load
conditions [35].

A+
IGBT gate
Drive
Optocoupler

B+

io

IGBT gate
Drive
Optocoupler

VAB
A

Load

Vdc

A-

B-

IGBT gate
Drive
Optocoupler

IGBT gate
Drive
Optocoupler

Controller
Figure 3.1 Block Diagram of Inverter System Including Controller

3.4 Inverter Loss Components based on Loss Model

Evaluating the losses associated with the inverter provides a clearer idea of the reasons behind the
reduction of the inverter efficiency at low power levels. Here a detailed mathematical analysis of
the losses in the SPWM inverter is made. Based on these equations for estimating the various

28

losses in the inverter, efficiency at light load and heavy load while varying the switching frequency
from 200Hz up to 20kHz can be calculated.

Typically, two thirds of the power loss in a hard switched inverter is the result of conduction and
switching losses in the inverter devices [32], [36]. Conduction losses occur due to the on-state
voltage across the device as well as the current flow through the device while it is conducting
current. More precisely, conduction losses occur between the end of the turn-on transition and the
beginning of the turn-off transition [36]. An effective model of conduction losses includes the
effect of device on-voltage and conduction resistance [37]. Switching losses arise from the
transient situation where both device voltage and current are changing as the device is turning on
or turning off [38], [39]. Evaluation of the conduction and switching losses can be done using
simplified device models described in [40], [41] and [42]. However, there are also other losses
associated with inverter operation. These include gate driver circuit loss, control circuit loss and
losses due to snubber circuits.

The gate drive circuit losses have been calculated but it is observed that they contribute a very
small amount of the total loss. The snubber circuit loss and the control circuit losses have been
ignored while developing the overall loss model given that they contribute to a very minimal
amount of loss compared to switching loss and conduction loss.

29

3.4.1 Conduction Losses

To evaluate the conduction loss through a simplified model which is appropriate for both IGBT
and diodes, the device is simplified as a constant voltage drop in series with a linear resistor [42].
The on-state voltage of an IGBT and a diode can be calculated using an IGBT datasheet. During
the time that the IGBT is on, the collector to emitter voltage, , is given by

= +

(3.1)

The same approximation can be used for the anti-parallel diode, giving

= +

(3.2)

Here, voltage source represents the IGBT on-state zero-current collector-emitter voltage and
stands for collector to emitter on-state resistance. Similarly, denotes on-state zero-instantaneous
current forward voltage for the antiparallel diode and stands for diode on-state resistance.
and are the currents flowing through the IGBT and diode respectively. The parameters , ,
and can be estimated directly from the component datasheets [40].

Figure 3.2 to 3.5 show the derivation of the parameters using datasheet information. Note in [41]
that a log-log plot of current vs. voltage is used to obtain the parameters, but is it suggested here
that linear scales provide more accurate parameter estimation related to the expected operating
30

conditions of the inverter transistors. As mentioned before, the rated RMS output current is 1.136A
and for a 4.5% load the corresponding output current is 62.9mA. Figure 3.2 and Figure 3.3 show
the derivation of the , , and parameters when the transistor (IGBT) is operating under
rated load conditions. Figure 3.4 and Figure 3.5 show the derivation of the , , and
parameters when the transistor is operating under light load conditions.

Figure 3.2 IGBT and (, =

) at Rated Load Condition; note that points

plotted here are taken from the Device Datasheet

Figure 3.3 Diode and (where, =

) at Rated Load Condition; note that points

plotted here are taken from the Device Datasheet


31

Figure 3.4 IGBT and (, =

) at Light Load Condition; note that points

plotted here are taken from the Device Datasheet

Figure 3.5 Diode and (where, =

) at Rated Load Condition; note that points

plotted here are taken from the Device Datasheet


32

To simplify the calculation of the device average and RMS currents, the load current is assumed
to be sinusoidal. Power dissipated in a component with a constant voltage drop is the average
current times the voltage drop [40]. The RMS current squared times the resistance signifies the
power dissipated in a resistor. The average and RMS currents of the IGBT and diode in an inverter
(given sinusoidal pulse width modulation) are [40]

= () [2 +

cos
8

() = ()8 +

= () [2

cos

(3.4)

cos
8

(3.3)

() = ()8

(3.5)

cos

(3.6)

Here, () denotes the peak load current, denotes the load power factor angle, the
modulation index, , denote the average currents and () , () denote RMS currents
flowing through the IGBT and the antiparallel diode [40]. The conduction losses in the IGBT,
and diode, are obtained using [40], [41].

= + () 2

(3.7)
33

= () 2

(3.8)

The total conduction losses, of the four IGBTs along with their anti-parallel diodes can
be calculated from

= 4( + )

(3.9)

The total conduction loss associated with the inverter is found easily using equation (3.9). Equation
(3.9) shows that the conduction losses depend on the load conditions [41], [42], [43], [44].

3.4.2 Switching Losses

In power inverters, switching loss typically contributes significantly to the total system losses. The
switching loss in the IGBT depends on the IGBT and diode's dynamic characteristics [45]. Three
components of the switching losses in the hard switching inverter can be identified: IGBT turn on
losses, IGBT turn off losses, and the losses due to diode reverse recovery. During turn-on, the
semiconductor is exposed to a high current peak as a consequence of the reverse recovery of the
freewheeling diode. At the same time the collector-emitter voltage is still high, thus causing high
switching losses. During turn-off, the losses can be even higher due to the long collector current
tail [44]. So, the turn-on losses are due to the rate of current change and the stored charge in the
free wheel diode. On the other hand, the turn-off losses depend on the speed of the gate drive and
the IGBT's current tail due to the recombination of minority carriers [46]. The semiconductor is
34

said to be hard switching under these conditions of simultaneous high current and high voltage
during the switching transient.

Evaluation of the switching losses, in the hard switching inverter consisting of four 15A, 600V
IGBT and ultrafast soft recovery diodes, can be done using the measured values of switching
energy from the data sheets. Generally, datasheets provide the values of turn-on and turn-off
energy ( and ) for a conventional test voltage and current ( and ). The calculated
values of turn-on energy comprise the losses due to diode reverse recovery and tail current losses.
The total energy loss during turn on and turn off transients of the switch, , can be calculated
based on [40] using

()

= ( + )

(3.10)

Equation (3.10) represents as the bus voltage, () as the peak load current and as the
correction factor to account for the gate drive impedance. The total switching losses, , of
the proposed inverter can found using

= 4

(3.11)

Here, denotes the SPWM switching frequency [43, 44]. Evidently, the switching losses in the
hard switching inverter are directly proportional to the switching frequency. Further, from equation
(3.11) the switching energy is proportional to the voltage across the device during switching, so
35

the losses can be eliminated if the voltage across the device is zero during the switching. This kind
of switching technique is called soft-switching or zero voltage switching (described in section 2.6)
but the added components will result in cost and reliability penalties [47].

3.4.3 IGBT Gate Drive Losses

IGBTs are voltage controlled devices and require a gate voltage to establish conduction between
collector and emitter. To provide the gate voltage, the FOD3184-ND is used as the gate drive
optocoupler. The total gate driver power loss can be derived from the summation of the total power
dissipated for the emitter ( ), internal circuitry ( ) and the output ( ) of the
IGBT driver IC [48]:

() = + +

(3.12)

IGBT total gate capacitance, , is the total gate charge divided by the gate drive supply
voltage [48].

(3.13)

This means that the charging and discharging the IGBT gate is equivalent to the charging and
discharging a capacitor. Hence the power dissipated for the output of the IGBT driver IC can be
defined by
36

= 2

(3.14)

Where, is the switching frequency. The power dissipated in the IGBT driver emitter can be
derived from the diode forward current , maximum diode duty cycle D and diode forward voltage
[48]:

(3.15)

Finally, the power dissipated in the IGBT driver internal circuitry depends on , the collector
current, and the collector to emitter voltage ( ). Note the collector to emitter voltage can
be any value between a minimum of -0.5V and a maximum of the device peak forward rating.
Thus

= ( )

(3.16)

3.5 Inverter Efficiency Calculation

Using equation (3.9) for conduction loss, equation (3.11) for switching loss and equation (3.12)
for gate drive loss, the different loss components of the proposed inverter at different load and
switching frequency conditions are calculated. Shown in Table 3.1 are the results for the
proposed inverter system (the topology is shown in Figure 3.1, and system details are provided in
Chapter 4). The table shows a decrease in switching and drive circuit loss for each load condition
37

when the switching frequency is decreased. Whereas, as shown in Table 3.2 the improvement in
the inverter efficiency in relation to the decrease in the switching frequency suggests that
reduction of switching frequency results in significant inverter efficiency improvement
especially at low power. The results of the proposed inverter loss model is presented for the full
voltage design and also for the low voltage scaled down design.

Table 3.1 Power Loss Calculation under Different Load Conditions for Variable Switching
Frequency (for 388.96, Rated Power 200W)

(kHz)

()

()

()

20

2.376

12.478

0.120

Full Load

10

2.376

6.238

0.114

(200W)

2.5

2.376

1.560

0.110

0.20

2.376

0.124

0.108

20

0.075

2.159

0.120

Light Load

10

0.075

1.079

0.114

(9W)

2.5

0.075

0.269

0.109

0.20

0.075

0.216

0.108

38

Table 3.2 Efficiency under Different Load Conditions for Variable Switching Frequency
(for 388.96, Rated Power 200W)

(kHz)

()

()

()

(%)

20

200

14.974

214.974

93.03

Full Load

10

200

8.728

208.728

95.82

(200W)

2.5

200

4.046

204.046

98.02

0.20

200

2.608

202.608

98.71

20

2.354

11.354

79.27

10

1.268

10.268

87.65

2.5

0.453

9.453

95.21

0.20

0.399

9.399

95.75

Efficiency

Light
Load
(9W)

The efficiency () calculated above is the reference for simulation and experimental
implementation of the proposed inverter. Ideally, inverter output power should be equal to the
inverter input power given that there is no loss in the ideal inverter system. Practically, there are
always some losses in the inverter system and input power can be expressed as sum of output
power and total loss. The basic formula for calculating inverter efficiency is, =
/ . Considering inverter loss explicitly, the formula can be re-written
as, = / ( + ). For example for 2.5kHz, the output
39

power is 200W, the total losses are 4.046W so the input power is now, 204.046W, thus the
efficiency is, = 200/ (200 + 4.046) = 98.02%. The efficiency calculations in Table 3.4 in the
last column is also done in the same way as with the sample calculation for 2.5kHz.

The trend of an increase in inverter efficiency with decreasing switching frequency is evident from
the efficiency calculations (Table 3.2). For example, inverter efficiency for 9W is 79.27% when
operated at 20kHz and 95.75% when operated at 200Hz. Interestingly the inverter efficiency does
not change dramatically as the switching frequency changes from 2.5kHz to 200HZ (i.e. 95.21%
to 95.75% for 200W in Table 3.2). The inverter loss model analysis based on different loss
components is done for an inverter circuit operating with a sinusoidal output current (i.e. no
harmonic distortion). Harmonic distortion, spikes, and controller losses are circuit specific
characteristics and cannot be modeled accurately without observing implemented inverter circuit
characteristics. To obtain the values of the equation parameters ( , , , in Equation. 3.1,
3.2), the datasheet graphs (log-log scale) have been linearized (Figure 3.2, 3.3, 3.4, 3.5) to
approximate required parameters.

Since a scaled down version (12.856W) of the full voltage design (200W) was implemented
experimentally, a similar kind of loss model analysis and efficiency calculation is performed for
the scaled down version as shown in Table 3.3 and Table 3.4.

40

Table 3.3 Power Loss Calculation under Different Load Conditions for Variable Switching
Frequency (for 25, Rated Power 12.856W)

()

()

()

()

20

2.375

1.248

0.120

10

2.375

0.624

0.114

2.375

0.312

0.111

2.5

2.375

0.156

0.110

0.20

2.375

0.013

0.108

20

0.074

0.889

0.120

10

0.074

0.444

0.114

0.074

0.223

0.111

2.5

0.074

0.111

0.110

0.20

0.074

0.009

0.108

Full Load
(12.856W)

(0.578W)

41

Table 3.4 Efficiency under Different Load Conditions for Variable Switching Frequency
(for 25, Rated Power 12.856W)

()

()

()

()

(%)

20

12.856

3.741

16.597

77.46

10

12.856

3.111

15.967

80.52

12.856

2.796

15.652

82.14

2.5

12.856

2.639

15.495

82.97

0.20

12.856

2.494

15.350

83.75

20

0.578

1.083

1.661

34.80

10

0.578

0.632

1.210

47.77

0.578

0.408

0.986

58.62

2.5

0.578

0.295

0.873

66.20

0.20

0.578

0.191

0.769

75.16

Efficiency

Full Load
(12.856W)

Light Load
(0.578W)

Similar to the full voltage design, Table 3.4 shows an increase in efficiency as the switching
frequency is decreased for the scaled down design. For example, inverter efficiency for 0.578W is
34.80% when operated at 20kHz and 75.16% when operated at 200Hz.

42

The results found from the mathematical analysis above, show that as the switching frequency
decreases, the losses in the inverter circuit decrease. As seen, this can be attributed to the change
in switching losses, unlike conduction and drive losses that remain nearly constant for a given load
condition.

3.6 Chapter Summary

A detailed mathematical analysis of the losses in the inverter circuit is presented in this chapter.
An inverter loss model is applied to show how reducing switching frequency improves the inverter
efficiency. The loss model gives information about how the transistor switching losses impact
inverter efficiency when the power level drops.

43

Chapter Four: Simulation Results for Proposed Full Bridge Inverter

4.1 Schematic of the Full Bridge Inverter System under Study

Before going to a practical circuit implementation, the simulation of the inverter circuit is
performed in OrCAD PSpice. PSpice is a computer simulation program that models the behavior
of an electrical circuit containing analog devices. PSpice is often chosen by design engineers for
its ability to simulate practical load characteristics. It is essentially a computer based breadboard
which allows prediction of AC and DC steady state waveforms and to perform transient analysis.
The simulation results from the PSpice inverter circuit model can be compared with the
mathematical model and with the experimental data. By verifying the PSpice model against
experimental data and the mathematical model, parametric studies of the inverter efficiency at
reduced switching frequency at low power may be conducted confidently. In PSpice, the practical
model of the IGBT can be used for the hardware circuit analysis and sinusoidal pulse width
modulation (SPWM) can be simulated in the control. Using the practical model of IGBT
(IRG4BC20UD) is useful and significant to observe the impact of switching loss associated with
switching frequency on the inverter efficiency. Also, the equivalent simplified PSpice models of
the heavy load and light load conditions help to predict the overall efficiency outcome of the
simulated inverter before going to hardware circuit analysis. To observe the impact of decreasing
switching frequency on the inverter efficiency, switching frequency is varied over a range of
20kHz to as low as 200Hz. This wide frequency range is useful to observe the change in inverter
efficiency as one goes from higher to lower frequency for heavy load and light load conditions.
The inverter system under study is shown in Figure 3.1 in Chapter 3 and also in Appendix A.
44

4.2 Simulation Circuit Description

The inverter circuit simulated in PSpice is a Voltage Source SPWM inverter. The simplest dc
voltage source for Voltage Source Inverter (VSI) consists of a battery bank [49]. The battery bank
usually consists of several batteries in series and/or parallel combination. Solar Photovoltaic cells
can also be employed to create the dc voltage source [50]. The proposed inverter is simulated in
PSpice for the full voltage design and also simulated for the low voltage scaled down design.

4.2.1 SPWM Controller

The number of pulses per half cycle is typically on the order of 100 pulses. As discussed earlier in
section 2.5, because of the advantages specific to the proposed inverter, unipolar SPWM is selected
as the suitable switching technique. Choosing the optimum switching technique for the IGBT
switching is an important design consideration for the proposed inverter.

The PSpice schematic shown in Figure 4.1 generates SPWM control signals. The sinusoidal signal
generator produces, a sinusoidal pulse signal at a frequency of . The control
frequency, is equal to the output frequency (i.e. 50HZ) of the output voltage [39].
The modulation index is set to be 0.8. The control signal, is then rectified through a
precision rectifier in order to implement unipolar SPWM topology (discussed in detail in section
2.4.1.2).

45

IF(V(%IN1)-V(%IN2)>0, 1, 0)

g1

1-V(%IN)
Inverter

g4

ABS

Comparator

g3
+

1-V(%IN)
Inverter

+
+

_ Vcontrol

Vx

1-V(%IN)
Inverter

g2

Vy

Vcarrier

Figure 4.1 Schematic of Controller for Generating SPWM Gating Signals as used in PSpice
Simulations

The triangular wave generator generates the carrier signal at the switching frequency of
. The switching frequency, = 2 , where N is the number of switching angles per quarter
cycle [50]. The switching frequency controls the speed at which the inverter switches are
turned on and turned off. In this thesis, the switching frequency is varied in a range of 20kHz to
200Hz. The change in switching frequency is made by changing the value of N as the output
frequency, is fixed at 50Hz.

Once the two signals are generated ( and ) the four gating pulses are generated by
comparing the triangular signal, with the absolute of the sinusoidal signal, [51].
The four gating signals as shown in Figure 4.2 is then applied to the four IGBT gates of the full
bridge inverter in order to create desired AC signal at the inverter output.

46

Figure 4.2 Simulated Waveforms for SPWM with Unipolar voltage switching (a) Sinusoidal
Reference waveform and Triangular Carrier waveform, (b), (c),(d),(e) gating pulses for
+ , , + and respectively after Sine and Triangular comparison (Total Time Interval
= 20ms, 1ms time division (horizontally), 5V voltage division (Vertically))

4.2.2 Gate Drive Circuit

It is customary to protect the IGBTs power devices used in the inverter so that they can continue
to function despite somewhat unpredictable conditions that characterize the renewable energy
scenarios. In particular, designers build in protection to avoid damage resulting from conditions
such as under voltage, over voltage, short circuits etc. [52].

An optocoupler is often used in IGBT gate drive circuitry to isolate the controller part from the
rest of the circuit. The optocoupler output diode is connected to a totem-pole pair of BJT devices
as shown in Figure B-3 (Appendix B) to provide fast switching of the IGBT (note that an isolated
47

supply is needed for the totem-pole pair). In a PSpice simulation the optocoupler and totem-pole
arrangement is modelled using a Gain Circuit, also called an E Block, as shown in Figure 4.3 (a).
Within the PSpice Gain Circuit is an isolation op-amp circuit, illustrated in Figure 4.3 (b). Shown
in Figure 4.3 (c) is a model of the internal op-amp structure. The main characteristics of an op amp
are very high input resistance (1 ), very low output resistance ( ) and very high gain (A is on
the order of 105). Note that a gate drive circuit must be able to provide a voltage of 15V to 20V
between gate and source of the IGBT, to ensure that the IGBT is fully on and in a state of low
conductance [53].

Vn
R1

Ro
|+

Vo
Vp

Ro

_
Vn

Vo

A(Vp-Vn)

Vn

Vp

(a)

R2
Vp

(b)

+
_

++
__

Vo

A(Vp-Vn)

(c)

Figure 4.3 PSpice Modelling of IGBT Gate Drive Circuit. (a) PSpice Gain Circuit which is
also called an E Block (b) Op Amp Symbol (c) Op Amp model.

4.2.3 IGBT Switching Circuit

In order to generate a modified sine wave at the inverter output requires both a positive and
negative voltage across the load, for the positive and negative parts of the wave respectively [54].
A modified sine wave as shown in Figure 4.4 at the output of the inverter can be achieved from a
48

single source through the use of four IGBT switches arranged in an H-Bridge configuration as
shown in Figure 4.1. In order to minimize power loss and utilize higher switching speeds, NChannel IGBTs were chosen as switches in the H-bridge inverter circuit.

Figure 4.4 Simulated Output Voltage Waveform of the Inverter at 2.5kHz for SPWM with
the Unipolar Switching (Total Time Interval = 20ms, 1ms Time Division (Horizontally),
50V Voltage Division (Vertically))

The top two IGBT switches (IRG4BC20UD) are + and while the bottom switches are + and
(shown in Figure 3.1) each co-packaged with HEXFREDTM ultrafast, ultra- soft-recovery
anti-parallel diodes for the use in H-bridge configurations [Appendix B]. The anti-parallel diodes
provide an alternate path for the load current if any of the power switches are turned off. For
example, if the lower IGBT ( ) in the left leg is conducting and carrying current towards the
negative dc bus, this current would regulate or commutate into the diode across the upper IGBT (
+ ) of the left. In order to avoid a short circuit of the DC bus, both IGBTs of the same leg can
never be conducting at the same time. In the unipolar SPWM switching scheme the output voltage
49

of the inverter swings from 0 to + and 0 to as shown in Figure 4.5. In the proposed
inverter switching scheme, when the switches + and are kept on, the output voltage across the
load is equal to + (311V). When and + are turned on, then at that time the output voltage
is equal to (-311V) [19]. The logic behind the switching of the devices in the leg connected
to and is given in Table 4.1,

Table 4.1 Unipolar SPWM Switching Logic


Switching Logic

Switch On

Voltage

State
>

= +(

<

= -(

>

= +(

<

= -(

Here, is denoted as the voltage of the reference signal in negative half cycle [19].

The operating principle of the full bridge inverter is similar to the half bridge inverter except a half
bridge inverter has only one switching device so that only the positive part of the sine wave can
get through the inverter. On the contrary, the single phase full bridge inverter has four switching
devices but instead of just clipping off half the wave it reverses the polarity of half the wave,
thereby increasing the efficiency and doubling the frequency [33].
50

4.2 Different Load Conditions

At rated power, in a rural typical household, loads such as a table fan, a refrigerator, a sewing
machine and lights could all be operating at the same time. However, there are situations where
the total load can be quite light, for example, at night only a single CFL (compact florescent lamp)
and/or a single cell phone charger.

To represent light loads, an equivalent model of the Compact Fluorescent Lamp (CFL) has been
simulated in PSpice. Typically, CFLs have two main components: a magnetic or electronic ballast
and a gas-filled tube (also called bulb or burner). Modern electronic ballasts contain a small circuit
board with rectifiers, a filter capacitor and usually two switching transistors [55], [56]. The
incoming AC current is first rectified to DC, then converted to high frequency AC by the
transistors, connected as a resonant series DC to AC inverter. The resulting high frequency is
applied to the lamp tube [57]. In the proposed inverter, only the ballast stage of the CFL consisting
of a rectifier circuit and a parallel RC circuit is simulated in PSpice. The simplified equivalent
PSpice model as shown in Figure 4.5 is useful for observing the nonlinear behavior [58], [59] of
the 9W CFL as shown in Figure 4.6 and Figure 4.7. In particular, the harmonic limits applied to
the CFLs are less stringent compared to LED (Light Emitting Diode) lighting [60], [61], [6] and
[7]. CFL lighting, even with high THD (<100%), is still used in rural locations [62].

51

D1
D1N3940

Cdc
0.7 uF

D3
D1N3940

Rload
5.315k

Inverter
Block
Vsystem =
220VRMS
Frequency= 50hz

D2
D1N3940

D4
D1N3940

Figure 4.5 Schematic of Equivalent Circuit of CFL Model connected as a load at the
inverter output

As shown in Figure 4.6, the passive front-end single-phase diode bridge rectifier is the standard
circuit for converting AC to DC. The rectifier stage in a CFL circuit has four rectifier diodes. In
this case, D1N3940 is used in the rectifier circuit which is commonly used in full bridge rectifier
circuits for CFL for its low forward voltage of 1V for a maximum forward current of 10 mA. This
circuit operates by first rectifying the single-phase input ac system voltage, (e.g. 220V
RMS), or rectifying the voltage delivered at the output of the inverter circuit, to produce a positive
cycle voltage waveform, , at twice the system frequency (i.e. 100Hz compared to system
frequency 50Hz). This voltage is then applied to the load capacitor, . When is greater than
, the load capacitor will start to charge and draw current from the inverter. At this point, as
there is no inductance present in the charging circuit, the dc current of the charging circuit
would cease instantaneously. During the time it takes to reach zero, will continue to
increase. Thus, the charging and discharging process results in non-linear operation of the bridge
rectifier circuit [60]. The diode bridge rectifier at the front end of the electronic ballast influences
52

that the CFLs draw non-linear current from the supply [60]. The charging state of the CFL load
circuit can be represented by a RC circuit driven by the rectified system voltage, [61].

In Chapter 4, the light load is considered to be a 9W CFL and the heavy load is considered to be a
load model of 200W combined heavy loads. However, in the experimental setup, the output power
for both cases (heavy and light) is scaled down to 12.856W and 0.578W respectively for a 25
supply input for practical reasons (discussed in detail in section 3.2).

The CFL model parameters were chosen in a way so that the power rating of the CFL can be
maintained at 9W. The set of equations to calculate the load circuit parameters is shown in section
5.3 of Chapter 5 for the scaled down version. The full voltage design parameters are also calculated
in the same manner using the same set of equations except the supply voltage is different than the
scaled down version. After connecting the equivalent CFL model as a load at the inverter output,
the current and voltage waveform as shown in Figure 4.6 and Figure 4.7 was observed.

53

Figure 4.6 Simulated Current Waveform from the Equivalent Circuit of CFL Model
connected as a Load at the Inverter output (Total Time Interval = 30ms, 1ms time division
(horizontally), 10mA current division (Vertically))

Figure 4.7 Simulated Voltage Waveform from the Equivalent Circuit of CFL Model
connected as a load at the inverter output (Total Time Interval = 20ms, 1ms time division
(horizontally), 50V voltage division (Vertically))

To observe the inverter efficiency at rated power (200W) along with the change in frequency an
equivalent model of the heavy load is simulated in PSpice. In the heavy load simulation an
inductive-resistive load as shown in Figure 4.8 was used.
54

Rload
Inverter
Block
Vsystem = 220VRMS
fSystem = 50hz

146.182

L1
0.3508 H

Figure 4.8 Schematic of Equivalent Circuit of CFL Model connected as a load at the
inverter output

When the inverter is connected to an inductive-resistive load, the anti-parallel diode of the IGBT
helps the IGBT to perform as a fully functional switch. The antiparallel diode connected to the
IGBT provides suitable reverse current for the inductive load. The current and voltage waveform
shown in Figure 4.9 and Figure 4.10 represent typical heavy load characteristics.

Figure 4.9 Simulated Current Waveform from the Equivalent Circuit of Heavy Load
Model connected as a load at the inverter output (Total Time Interval = 30ms, 1ms time
division (horizontally), 1A current division (Vertically))
55

Note that, as shown in the above Figure 4.10, the inverter output current waveform operates (heavy
load condition) in a transient state during approximately the first 10ms and reaches almost steady
state after 10ms of its operation period (the LR time constant of the load is 2.40ms).

Figure 4.10 Simulated Voltage Waveform from the Equivalent Circuit of Heavy Load
Model connected as a load at the inverter output (Total Time Interval = 20ms, 1ms time
division (horizontally), 50V voltage division (Vertically))

4.3 Simulation Results

To observe the trend of improved inverter efficiency along with decrease in switching frequency,
the proposed inverter has been simulated for the full voltage design and also for the scaled down
low voltage design (discussed in detail in section 3.2). For both cases (full voltage design and the
scaled down version) simulation results of the PSpice inverter under different load conditions and
variable switching frequencies are presented in Table 4.2 and Table 4.3.
56

Table 4.2 Simulation Results Under Different Load Conditions For Variable Switching
Frequency (For 388.75)
Switching

Output

Total

Input

Efficiency

Frequency

Power

Loss

Power

()

()

()

()

(%)

20

200

35.182

235.182

85.04

Full Load

10

200

22.281

222.281

89.98

(200W)

2.5

200

17.372

217.372

92.01

0.2

200

12.669

212.669

94.04

20

5.072

14.072

63.96

Light Load

10

2.632

11.632

77.37

(9W)

2.5

2.492

11.492

78.32

0.2

1.388

10.388

86.64

Load

The efficiency calculated shows an increasing trend as the switching frequency decreases from
20kHz to 200Hz. The change in efficiency is more evident at the light load scenario when only a
9W CFL is connected at the inverter output. As shown in the fourth column of the above table,
Table 4.2, the efficiency is 63.96% for 20kHz and the efficiency for the given constant load
changes into 86.64% when the inverter is operating at a switching frequency of 200Hz. The
efficiency of the simulated inverter is calculated using the same efficiency equation (Efficiency =
57

Output Power/ Input Power, where, Input Power = Output Power+ Total Loss) described in section
3.5 of Chapter 3.

Table 4.3 Simulation Results Under Different Load Conditions For Variable Switching
Frequency (For 25)
Switching

Output

Total

Input

Efficiency

Frequency

Power

Loss

Power

()

()

()

()

(%)

20

12.856

7.116

19.972

64.37

10

12.856

6.341

19.197

66.97

12.856

5.262

18.118

70.96

2.5

12.856

3.841

16.697

76.99

0.20

12.856

3.546

16.402

78.38

20

0.578

1.112

1.690

34.20

10

0.578

0.681

1.259

45.91

0.578

0.450

1.028

56.22

2.5

0.578

0.386

0.964

59.96

0.20

0.578

0.298

0.876

65.98

Load

Heavy Load
(12.856W)

Light Load
(0.578W)

The efficiency calculations as shown above show an increasing trend even for a very low power
level of 0.578W (from 34% to 65.98%). As shown in Table 4.3, for both cases (heavy load and
58

light load) an additional switching frequency of 5kHz is added between the switching frequencies
10kHz and 2.5kHz. The significance of adding an additional switching frequency in the efficiency
calculation table is that, while implementing the experimental inverter prototype, the switching
frequencies were maintained within a close range, where the highest switching frequency was
5kHz and lowest was 200Hz with a 2.5kHz switching frequency in between. Thus, while
comparing the mathematical analysis, simulation and experimental efficiency results of the
inverter, the comparison can be done for all the three switching frequency levels (e.g. 5kHz,
2.5kHz, 200kHz). The reason behind keeping the switching frequencies within a close range for
the scaled down version from 5kHz to 200Hz is that the same snubber circuit was used for varying
switching frequencies for practical purposes (c.f. section 5.2.6). The reason for ignoring the higher
switching frequencies (e.g. 10kHz and 20KHz) is that at very low power (e.g. 0.578W) where the
output current is very small (e.g. 62.9mA RMS), the current waveform becomes excessively
distorted by the effect of electromagnetic interference caused by the higher switching frequencies.
The result was overheating of the transistors. In order to correct this problem, the snubber would
have needed to be re-designed which is not done in practice (the snubber components are usually
fixed values).

A Fourier analysis of the output current is performed in PSpice for both full load and light load
conditions. Based on the analysis, Total Harmonic Distortion (THD) is observed for both
conditions. For full load it is observed that the output current is a near-sinusoidal waveform with
total harmonic distortion of around 17.50% for the lowest switching frequency of 200Hz. The
THD of the output current at 200Hz which is 17.50% is within the acceptable range of 20% for
household appliances for stand-alone solar home systems [6]. When considering a CFL as a load
59

for the inverter, the output current has a higher total harmonic distortion of 57.14% for a switching
frequency of 200Hz which is also acceptable (<100%) for non-linear loads like a CFL which
produces a high current THD factor [35], [61], [62], [63] and [64].

4.4 Chapter Summary

Presented in this chapter is the inverter simulation model. The model consists of a dc power supply,
controller circuit, gate drive circuit, switching circuit implemented for heavy and light load
conditions. Simulation results support the results from the mathematical analysis of the previous
chapter. Hence, suggests that reducing the switching frequency at low power decreases the
switching loss, thus increasing efficiency for the same load. The THD of the output current of the
inverter under heavy load and light load meet the stand-alone solar home THD requirements [6],
[35]. Hardware implementation can be conducted confidently based on the design considerations
and analysis of results obtained from the PSpice simulation of the proposed inverter circuit. The
gate drive circuit losses did not reduce the inverter efficiency significantly as was found in the
mathematical analysis. The results found from the loss model developed in Chapter 3 and the
results obtained from the simulations shows agreement in the efficiency trend. The mathematical
loss model is useful to observe the increasing efficiency trend but it is extremely important to
simulate the inverter through PSpice before going to practical implementation. The PSpice
simulation platform is useful to verify that the individual components work well together in order
to provide useful AC at the inverter output.

60

Chapter Five: Experimental Results

5.1 Introduction

The experimental verification of the proposed single phase full bridge SPWM inverter to improve
inverter efficiency at low power is described and discussed in this chapter. A prototype of the
inverter circuit was built on a breadboard which is then controlled by the MSP-430 microcontroller
that includes a PC interface for programming purposes. This chapter has two main parts. Section
5.2 to 5.3 provide a description of the hardware components and the process to select component
values. These components are: the control circuit for generating switching signals, the isolation
circuit, the gate drive circuit, and the snubber circuit. Calculations are provided for snubber
component values, load component values and current sense resistor values. In the second part of
this chapter, the experimental results are presented, which are compared with the simulation results
of Chapter 4 as well as the mathematical results of Chapter 3. The 25 dual mode DC power
supply voltage was applied to the inverter in order to simplify the experimental work instead of
high voltage (389.98). The mathematical loss model in Chapter 3 and the PSpice model in
Chapter 4, discusses both the full voltage and the scaled down versions. However, this chapter
only includes the low voltage design (i.e. the scaled down version at 12.856W rated power) which
was implemented experimentally in the lab set-up.

61

5.2 Hardware Overview

Figure 5.1 is a partial circuit diagram of the inverter (only the gating lines are shown of the
MSP430 microcontroller). The dual mode DC power supply provides a constant 5 voltage to
the DC-DC converters. Each DC-DC converter in turn provides a constant 15V supply to each
optocoupler. There are four optocouplers in the inverter circuit and each of them needs an isolated
power supply otherwise transistor gate inputs become shorted. The four DC-DC converters provide
the isolated power supply to each of the optocoupler ICs which provides isolation between a
microprocessor gating line and the IGBT transistor gate. The output of the optocoupler is
connected to the IGBTs gate through the gate resistor. As mentioned in the previous section, the
25 dual mode power supply provides the input voltage for the full bridge inverter. This is the
input voltage for the experimental inverter circuit (in practice a much larger supply voltage, for
example, 388.96 is needed in order for the inverter to power 220 equipment). Although
the supply voltage is a scaled down it is expected that a similar efficiency trend will be observed
as would be the case for the full voltage design. Chapter 4 described the full voltage design for
generating a 220 voltage at the output of the inverter. In section 4.3 the efficiency calculations
showed that the inverter efficiency increases with decreasing switching frequency irrespective of
the load (heavy or light).

The following sections detail how each specific part is constructed and interacts with other parts.
The hardware design for this project was divided into four main blocks (a) Power supply for
transistor drive circuit; (b) Isolation circuit for each transistor; (c) Full bridge inverter circuit and
(d) Snubber circuit for each transistor.
62

The MSP430 was used to generate four gating pulses. The MSP430G2553 is an ultra-low power,
mixed signal microcontroller with built-in 16-bit timers. There are up to 24 I/O capacitive-touch
enabled pins, a versatile analog comparator and a 10-bit analog-to-digital (A/D) converter. Up to
seven SPWM signals can be generated with the Timer A of MSP430G2553 [Appendix C]. In
Figure 5.1 only four SPWM output pins 2.1, 2.2, 2.4, 2.5 of the MSP430 are shown which has
been used to generate four gating pulses [65]. The MSP430 code that is used to generate these four
gating pulses is discussed in detail in Appendix D.

63

DC Power Supply
5V
25V
_

B+

A+
35 ohm
1
4
2

0.1uF

DC-DC
Converter

COM

50 ohm

50 ohm
5.1uF
10 ohm
10 ohm
Snubber Circuit

0.1uF

90 ohm

Optocoupler

90 ohm

0.1uF

0.1uF

35 ohm
4

1
2

DC-DC
Converter

Optocoupler

Load

Snubber Circuit
B-

A-

50 ohm

5.1uF

DC-DC
Converter

0.1uF

50 ohm

10 ohm

10 ohm

0.1uF

90 ohm

0.1uF
0.1uF

35 ohm
4
3

1
2

DC-DC
Converter

Optocoupler

Optocoupler
90 ohm
2.1

2.4
MSP430

2.2

2.4

Figure 5.1 Partial Schematic of the Experimental Inverter System

The inverter circuit as shown in Figure 5.1 was designed to deliver AC power to the loads used in
the households, but is a scaled down version, operating at a reduced voltage to simplify the
experimental work. Although it was not possible to run 230 equipment with this scaled down
version it was possible to investigate the variation in efficiency at different switching frequencies.

64

5.2.1 Power Supply for Transistor Drive Circuit

The gate drive circuit requires 15 to operate. To provide this voltage, a DC-DC
converter circuit was used, The DC-DC converter was NME0515SC which is an isolated 1W
single output DC/DC converter [Appendix B]. The output current of the DC-DC converter is 66mA
with an input current of 250mA and a conversion efficiency of around 80%. Minimum load to
meet datasheet specification which is 10% of the full rated load (1W) across the specified input
voltage range was used. Lower than 10% minimum loading will result in an increase in the output
voltage, which may rise to typically double the specified output voltage if the output load falls to
less than 5%. As per the datasheet, a 0.1uF capacitor was connected in parallel to the converter to
remove the noise coming from it and thus protect the overall circuit.

5.2.2 Isolation Circuit

An isolation circuit is needed to protect the microprocessor from the overvoltage damage. The gate
to signal isolation for inverter switches is generally achieved by means of optical isolator circuits.
As shown in Figure 5.2, the FOD3184 was used as the gate drive optocoupler [Appendix B].
FOD3184 is suitable for high speed IGBT such as IRG4BC20UD which has been used as the
switching device in this experiment. The voltage and current supplied by the optocoupler is ideally
suited for driving the specific IGBT rated at 600V/6.5A. As shown in Figure 5.2 and in Figure B3 [Appendix B] FOD3184 consists of an Aluminum Gallium Arsenide (AlGaAs) light emitting
diode optically coupled to CMOS detector with PMOS and NMOS output power transistors
integrated circuit power stage. As mentioned earlier a 0.1uF bypass capacitor was required to be
65

connected between pins 5 and 8. This 0.1uf capacitor acts in effect like a low pass filter, adding
some smoothing to the input signal and bypassing sharp spikes.

5.2.3 Optocoupler Circuit Operation

The input stage of the optocoupler IC is a light emitting diode (LED) that emits light when forward
biased. The light output of the LED falls on reverse biased junction of an optical diode as shown
in the optocoupler figure. The gate control pulses for the switch are applied to the input LED
through a current limiting resistor (R) of appropriate magnitude. These gate pulses, generated by
the MSP430 microcontroller, are essentially in the digital form. A high level of the gate signal is
an on command and a low level is connected to the ground point of the MSP430 microcontroller
ground. The anode is fed with the gating pulse generated by the MSP430. The circuit on the output
(photo-diode) side was connected to a floating dc power supply, as shown in Figure 5.2. The
control supply ground was isolated from the floating supply ground to the output. The circuit on
the output (photo-diode side is connected to a floating dc power supply ground of the output. In
the figure the two grounds have been shown by two different symbols. The schematic connection
in the figure indicates the magnitude of the reverse leakage current of the diode. When the input
signal is high, the LED conducts and the emitted light falls on the reverse biased p-n junction.
Irradiation causes generation of significant number of electron-hole pairs in the depletion region
of the reverse biased diode. As a result, the magnitude of reverse leakage current of the diode
increases appreciably. The resistor connected in series with the photo-diode now has higher
voltage drop due to the increased leakage current. The two transistors work together as a signal
comparator circuit and senses this condition. The signal comparator then outputs a high level
66

signal, which is amplified before being output. Thus an isolated and amplified gate signal is
obtained and may directly be connected to the gate terminal of the switch.

Power Supply
Vdd =15V(Floating)
NC

Input Pulse R

8
Photo- Diode

ANODE

CATHODE

NC

LED

Vdd
Vo2

Output
6

3
4

Control
Ground

Vo1

Vss

Floating
Ground

Figure 5.2 Internal Circuit of FOD3184 IC

In this thesis, FOD3184 optocoupler IC is employed to drive the IGBT. As shown in Figure 5.3
below the FOD3184 optocoupler IC isolates an IGBT from the controller circuit. Each optocoupler
requires a 15V isolated DC supply since an IGBT can only be placed in a low conduction state if
the gate to source voltage is much greater than the transistor threshold voltage. This 15 is
supplied by a DC to DC converter circuit (discussed in section 5.2.1). The optocouplers are
controlled by the MSP430 microcontroller unit.

67

MSP430 Microcontroller
Instructions

Oprocoupler IC

15V DC
From DC-DC Converter

IGBT Drive

IGBT

Figure 5.3 Block Diagram showing Optocoupler Interface

5.2.4 Full Bridge Inverter Circuit

The single phase full bridge switching circuit has been shown in Figure 5.1. Thus the single
phase full-bridge (often, simply called as bridge) circuit has two legs of switches, each leg
consisting of an upper switch (e.g. + , + ) and a lower switch (e.g. , ). Junction point of the
upper and lower switches is the output point of that particular leg. The emitters and bases of the
upper and the lower IGBTs are shorted as shown in the Figure 5.1. An IGBT is turned on and off
by the MSP430 microcontroller to produce a high frequency SPWM signal which is later converted
to a modified sine wave at the output [Appendix C].

The individual control signal for the switches needs to be provided across the gate and emitter
terminals of the particular switch. The gate control signals are low voltage signals referred to the
emitter terminal of the switch. For each IGBT switch, when gate to source voltage is more than
threshold voltage for turn-on, the switch turns on and when it is less than threshold voltage the
switch turns off. In this thesis, IGBT model IRG4BC20UD has been used for switching purposes.
The threshold voltage of the specific IGBT is +6V but the turn-on gate voltage magnitude is +15V
68

for quicker switching whereas turn-off gate voltage is 0V. The two switches of an inverter-leg are
controlled in a complementary manner as shown to generate the AC output of the desired
frequency (50 Hz) at the inverter output.

As shown in Figure 5.4 when the upper switch of any leg is on, the corresponding lower switch
remains off and vice-versa. When a switch is on its emitter and collector terminals are virtually
shorted. Thus with upper switch on, the emitter of the upper switch is at positive dc bus potential.
Similarly with lower switch on, the emitter of upper switch of that leg is virtually at the negative
dc bus potential. Emitters of all the lower switches are shorted with the collectors of the lower
switches as shown in Figure 5.1. Since gate control signals are applied with respect to the emitter
terminals of the switches, the gate voltages of all the upper switches must be floating with respect
to the dc bus line potentials. The emitters of the lower switches of both of the legs are at the same
potential (since they are connected to the negative DC bus) and hence the gate control signals of
lower switches need not be isolated among themselves.

However, there are spikes (with ringing effect) at the output current and output voltage waveform
of the inverter caused by the circuit parasitic inductance when a switch opens. To eliminate voltage
spikes a snubber circuit has been used which provides an alternate path to ground for the current
flowing through the circuit's parasitic inductance. The snubber reduces the voltage transient and
damps the subsequent ringing with the parasitic capacitance that occurs when the switch opens.

69

5.2.5 Snubber Circuit

Snubbers are circuits which are placed across the semiconductor devices (in this case IGBTs) for
protection and to improve performance [66]. They are voltage limiting or current limiting devices
connected to the power terminals (collector and emitter of an IGBT device). For voltage snubbing,
the snubber is connected in parallel with the power transistor to reduce the peak value of voltage
switching transients. Snubbers are used to prolong the life of contacts by reducing arcing in
semiconductor devices.

5.2.6 RC Snubber

There are many different kinds of snubbers. The two most common ones are the resistor capacitor
(RC) damping network and the resistor-capacitor-diode (RCD) turn-off snubber. Figure 5.4 shows
the RC snubber circuit is implemented in this thesis, to get rid of the spikes at the output voltage
of the inverter circuit [67].

An RC snubber is most commonly used in inverters for both rate-of-rise control and damping. The
RC snubber absorbs energy during each voltage transition and can reduce efficiency. Also, the RC
snubber can reduce the switching speed of the IGBT switch. Care must be taken in choosing the
value of resistor and capacitor to optimize the total performance [68]. In addition to removing
spikes at the inverter output signal, a RC snubber is also responsible to damp the parasitic ringing
in the circuit [69]. In these applications, the value of the resistor must be close to the characteristic
impedance of the parasitic resonant circuit it is intended to damp [70].
70

Rs
Cs
Snubber Circuit

Figure 5.4 Equivalent Circuit of the Snubber Circuit

If the DC supply is assumed to have negligible internal impedance, the worst-case peak current in
the snubber circuit can be calculated by the following equation [67],

(5.1)

where, = Snubber resistance, = Open circuit voltage

The energy dissipated by the snubber is the energy stored in the snubber capacitor, . Thus the
total energy dissipated through the resistor can be expressed as,

= 12 ( )2 2

(5.3)

= ( )2

(5.4)

where, = switching frequency, 2 = number of transitions per cycle


71

The snubber capacitance, has to meet two requirements. First, the energy stored in it must be
greater than the energy in the circuit's inductance.
1

2 > 2 2
2

(5.5)

which can be rearranged as,


> 2 / 2

(5.6)

where, = open circuit voltage, = closed circuit current and L = circuit inductance
Secondly, the time constant of the snubber circuit should be small compared to the shortest on time
expected, usually 10% of the on time.

<

(5.7)

10

where, = Shortest on-time expected

Using equations (5.1) to (5.6) as a guide [66], [67], the snubber resistor value was chosen to be
10 and the snubber capacitor to be 5.1uF for switching frequency 2.5kHz. The snubber reduces
ringing by limiting the peak voltage on a switching transistor (i.e. IGBT in this thesis). As shown
in equation (5.4), the snubber capacitance, is directly proportional to the switching frequency,
. So, according to the proportional relationship, should change every time the switching
frequency changes but in a lab set up it is impractical to change the snubber circuit of the inverter
prototype every time the switching frequency changes. value is calculated for 2.5kHZ switching
frequency and then the same snubber has been used for 2.5kHz and 5kHz as well. In order to use
72

the same snubber capacitor, the switching frequencies (200Hz, 2.5kHz and 5kHz) are kept in a
close proximity.

5.3 Load Modeling and Load Value Calculations

For the implementation of the scaled down version, the heavy load and light load are scaled down
to 12.856W and 0.587W respectively for a 25 supply input for practical reasons. In the
hardware implementation, the same CFL model parameters described in section 4.3 is converted
to match the scaled down power requirement. Figure 5.5 shows the CFL model with re-calculated
values for and . The equivalent CFL model used in Figure 5.5 is similar as the
simulation CFL model in section 4.2, except the sense resistor, is introduced in the load
circuit in order to measure the total current flowing into the load from the inverter output while
implementing in the hardware.

Idc
D1
D1N4148

Sense Resistor

Rsense
1.56

D3
D1N4148

ITotal

10.75
uF

Vrect
Inverter
Block
Vsystem = 14.14VRMS
Frequency= 50hz

CLoad

D2
D1N4148

Vdc
Rload
346.14

D4
D1N4148

Figure 5.5 Equivalent Circuit Model of CFL for Scaled Down Version

As shown in Figure 5.5, the passive front-end single-phase diode bridge rectifier is the standard
circuit for converting ac to dc in a CFL circuit which has four rectifier diodes. In this case,
73

D1N3940 is used at the rectifier circuit which is commonly used in full bridge rectifier circuits for
CFL for its low forward voltage of 1V for a maximum forward current of 10mA. The sense
resistance of the charging circuit, , represents the resistance through which the voltage drop
is measured to measure the total current, going through the load circuit of the inverter circuit
[55].

5.3.1 Determination of Component values for Light Load (CFL) Model

As shown in Figure 5.4 above, the load capacitor and the load resistor resembles a
RC parallel circuit. To obtain the values of load capacitor and load resistor a complete
analysis of the RC parallel circuit stage is discussed in this section [71], [72].

The total phase angle for the CFL circuit is given by,

= tan1(

(5.8)

where, is the total capacitive reactance in the parallel RC circuit and as shown in Figure 5.5
and is the load resistance.

The total impedance in parallel RC circuit can be expressed as,

(5.9)

2 + 2

74

which can be rewritten as follows,

= tan

(5.10)

Now combining equation (5.8) and equation (5.9) the following equation can be found,

Z=

2 tan
( tan )2 + 2

tan

(5.11)

tan 2 +1

Now, in order to find the value of the total impedance, of the RC parallel circuit, the value of
needs to be inserted. The value of is known as it comes from the load power factor,
cos . The power factor cos is a measure of the phase difference between the voltage and the
current. CFLs are known for their low power factor, which is usually the result of a significant
phase difference between the voltage and current at the load terminals, or it can be due to a high
harmonic content or a distorted current waveform. For 9W CFL, the power factor is usually 0.65
[73]. The main reason for a low power factor in a CFL as a load is that it draws current that is not
in phase with the voltage waveform. Poor power factor (<1) causes inefficiency in the delivery
of electricity to the end-user, requiring more energy to compensate for losses on the line but also
makes them an inexpensive and most available lighting option in rural households [74], [75]. In
this thesis although the light load power was scaled down to 0.578W but the power factor is still
maintained at 0.65 as with the full voltage design.

For a power factor of 0.65, it can be written that,


75

cos = 0.65

(5.12)

So, = cos1 (0.65) = 49.45

Here, tan 49.45 = 1.1669, so the total impedance of the RC parallel circuit can be derived from
the following equation,

1.169
(1.169)2 + 1

1.169
1.538

Therefore, = 0.76

(5.13)

Now, to find out the RMS current, , the total power can be written as

= cos

(5.14)

0.578 = 14.142 0.65


= 0.0629

(5.15)

The apparent power, S of the RC parallel circuit can be expressed as

=
Therefore,

= 14.1420.0629 = 0.8892VA

76

(5.16)

In other terms, apparent power can also be written as,


S=(

2
) Z

0.8896 =
So,

(14.142)2

Z = 224.879

(5.17)

Therefore by substituting the value of Z into equation (5.13) the load capacitance can be
found from the following equation

0.76

224.879 = 250

0.76

= 250224.879 = 10.75uF

And finally,
= tan
1.169

= 25010.75106
= 345.142

Only the input voltage is scaled down from 388.906V to 25V and the output voltage is scaled down
from 220 to 14.142 but the output current has been kept the same which is 62.9mA
, output current for light load condition.

The load capacitance can be calculated by using,


77

0.7 25
388.96

= 10

The load capacitor, = 0.7 for 9W load capacitor by following the above equations and
load resistor and capacitor calculations. The calculated values for 9W load capacitor and load
resistor is used in the equivalent light load circuit in PSpice circuit described in chapter 4. A scaling
25

factor of 388.96 was used to find out the approximate values of the load capacitor and load resistor
when the voltage is scaled down as mentioned before.

Similarly the resistance can be calculated as,


=

5.315103 25
388.96

= 341.616

5.3.2 Heavy Load Model Calculations

RLoad
Rsense
0.156
ITotal

LLoad

Vsystem = 14.144Vrms
fSystem = 50Hz

Figure 5.6 Equivalent Heavy Load for Scaled Down Version


78

As mentioned above, the heavy load is modeled as a simplified equivalent combination of a table
fan, mini refrigerator and a sewing machine. The heavy load is mostly inductive-resistive in nature.
When the inverter is connected to an inductive-resistive kind of load, the anti-parallel diode of the
IGBT helps the IGBT to perform as a fully functional switch [76]. The antiparallel diode connected
to the IGBT provides suitable reverse current for the inductive load.

To obtain the values of load resistor, and load inductance, in the context of heavy
load a complete analysis of the RL series circuit stage is discussed in this section.
The power factor for the heavy load is considered to be 0.8 [73], so the power factor angle can be
expressed as,

cos = 0.8 = cos1 (0.8) = 36.8699

(5.17)

The total phase angle can also be written as,

tan =

(5.18)

tan(36.8699) =
0.75 =

0.75 =
L=

0.75

Therefore, L = 0.00239

(5.19)
79

Now, the total power of the RL circuit can be expressed as,


P = cos
12.856 = 14.142 0.8
= 1.136 A

It is important to note that only the RMS voltage has been scaled down to 14.142V from 220V but
the current remains same for both power levels of 12.856W and 200W. Thus, the RMS current is
maintained at 1.136A.

By following Ohms law the total RMS current can also be expressed as,

(5.20)

The total impedance Z can be written as,

= 2 + (2)2

(5.21)

Substituting the equation (5.21) into equation (5.20),

2
+(2)2

(5.20)

By substituting the value of into equation (5.20),


80

1.136 =

14.142
2 +(2)2

Now, by substituting equation (5.19) into equation (5.20),


1.136 =
=
=

14.142
2 +(20.00239)2
14.142
2 {1+(20.00239)}2
14.142
1+0.753984

Therefore,
14.142

R = 1.1361.32438
= 9.9577

Substituting the value of resistance into equation (5.19), the load inductance,

(0.759.4012)
(250)

= 0.02379

Thus, the resistor is 9.9577 and the inductor is 0.02379.

81

5.3.3

Current Sense Resistors

As shown in Figure 5.6 the shunt resistor is used to measure current at the heavy load output.
Generally, when designing power supplies and regulated battery circuits, the aim is to eliminate
the risk of short circuits or over current conditions which are likely to damage other components.
Current resistors are a simple and economic means to protect the switch mode power supplies like
the inverter circuit. The current resistors, also called shunt resistors, are used to monitor the
current in a circuit and translate the current into a voltage that can be easily measured and
monitored. Such resistors have very low resistance values, typically less than 1 [77]. In this
thesis, for heavy load the sense resistor is taken to be 0.156 but for light load the sense resistor
was chosen to be 1.56. A bigger resistor for current measurement is required for this experiment,
as both the input and output current for light load is very small and difficult to measure accurately
with the available measurement tools.

The currents flowing into the inverter circuit which is the input current of the inverter and through
the load circuit which is the output current of the inverter is measure though the current resistors
for both heavy and light load cases but with different resistor values.

5.4 Experimental Results

The inverter prototype was developed and tested in a modular fashion. Several tests were
conducted to check if components are working individually. Then, components are integrated to

82

make the whole inverter circuit. After verifying inverter prototype operation, the results listed in
Table 5.1 were obtained.

5.4.1 Measurement Approach

For hardware result analysis the Tektronix 3100 oscilloscope data was used with RS232 serial data
transmission cable. The oscilloscope window was set to 20ms to capture a full cycle of the signal.
The data acquired through the RS232 was then transferred to the laptop through a Tektronix
software called Open Choice Desktop. The data is captured through waveform data capture and
then the data table is saved as a .csv file. The .csv file contains 10,000 samples for 20ms data. The
following data files are generated for each switching frequency using Open Choice Desktop:

1. Voltage at input and output side before the sense resistor.

2. Voltage at input and output side after the sense resistor.

3. Output voltage for specific load condition.

Once the voltage drop across the sense resistor is found, the voltage difference is divided by the
value of the sense resistor (0.156 for light load, 1.56 for heavy load) to find out the input and
output current. The voltage drop at the inverter output is taken by using the MATH function of the
oscilloscope. Input and output voltages are directly available from oscilloscope data. MATLAB
code was written for output and input power calculations. After acquiring the both voltage and
83

current vectors in MATLAB .csv file (10,000 samples within 20ms), they are integrated over time
using the trapz function. This quantity is further divided by 20ms to obtain the average input and
output power. After the input and output power was calculated the efficiency calculation is done
by dividing the output power by input power in a separate MATLAB file.

The values of hardware results (scaled down version) found through the data analysis of MATLAB
are shown in the Table 5.1.

5.4.2 Results and Discussion

The hardware results for a scaled down version of the full voltage design are shown in Table 5.1.

Table 5.1 Hardware Results of full load and light load

()

()

()

()

(%)

12.757

5.898

18.655

68

2.5

12.886

4.307

17.193

75

0.2

12.863

3.817

16.680

77

0.574

0.508

1.082

53

2.5

0.578

0.413

0.991

58

0.2

0.577

0.310

0.887

65

Full Load
(12.856W)

Light Load
(0.578W)

84

The results show the same trend as the simulations and loss model based efficiency calculation.
As mentioned before, only the scaled down version (12.856W rated power) has been implemented
in the lab. The inverter efficiency increases as the switching frequency is reduced. The hardware
results agree with the simulation and mathematical results of the scaled down version as shown in
previous chapters. The efficiency of the experimental inverter circuit as shown in the fourth
column in the above table (Table 5.1) is calculated by dividing the average output power by the
average
=20

=0

input

power.

The

average

output

power

is

given

by,

() ()
20, once the circuit is operating in steady-state, where the output

power is an integral of the output voltage times output current which is then divided by the total
time 20ms. The average input power is found in a similar manner, using, =
=20

=0

() ()
20, once the circuit is operating in steady-state, where the input

power is an integral of the input voltage times input current which is then divided by the total time
20ms.

The comparison between the results (scaled down version) of the mathematical analysis,
simulation model and the hardware prototype is presented in Table 5.2.

85

Table 5.2 Comparison of the Efficiencies


Switching

Mathematical

PSpice

Hardware

Frequency

Analysis

Simulation

Implementation

()

(%)

(%)

(%)

82.14

70.96

68

2.5

82.97

76.99

75

0.2

83.75

78.38

77

58.62

56.22

53

2.5

66.20

59.96

58

0.2

75.16

65.98

65

Load

Full Load
(12.856W)

(0.578W)

As shown in the above table, Table 5.2, there is a slight difference in efficiency between the
mathematical analysis and the other two models (simulation and hardware). One reason for the
discrepancy maybe the linearization performed using datasheet information (see Figures 3.2 to
3.5). However, the simulation is performed in PSpice using practical component models and thus
shows good agreement with the hardware test results.

5.5 Chapter Summary

This chapter discusses the details of how the inverter prototype was built and tested. The full
voltage design (rated power 200W) was not implemented in hardware. Instead a scaled down
86

design of 12.856W was used. The first part of this chapter describes the calculation of the
component values for the light and heavy load tests. The hardware results show the expected result
that reducing the switching frequency at low power decreases switching loss, thus increases
efficiency for the same load. The hardware results also show the similar results as simulation and
mathematical results as described in Chapters 3 and 4. The mathematical and simulation result
tables (Table 3.1 and Table 4.1) for the full voltage design shows less variation in efficiency than
at low power (Table 5.1).

87

Chapter Six: Conclusions

6.1 Summary

In Chapter 2, a detailed analysis of different techniques to improve inverter efficiency is carried


out. Five different techniques were discussed: hybrid switch (combination of MOSFET and
IGBT), enabling pulse skipping mode, implementing Zero Voltage Switching (ZVS) while turning
on and off, limiting current through burst mode LLC and using variable switching frequency in
discontinuous current mode (DCM). It was found that the decreasing switching frequency is
superior in improving the inverter efficiency at low power than the hybrid switch configuration,
ZVS and burst mode LLC. Also, there is a need for an optimum snubber circuit for each switching
frequency. For grid connected inverters, more complex control circuits for maintaining the power
quality at the inverter output are required. However, for a stand-alone rural solar home system the
power quality is traded off with the increased inverter efficiency at low power.

The research work is given in three main parts: Chapters 3, 4 and 5. Chapter 3 presents the
mathematical analysis of the proposed full bridge SPWM inverter and results found by using the
loss equations. Chapter 4 describes the PSpice simulation of the proposed inverter and the
simulation results for varying switching frequencies. In Chapter 5, the hardware implementation
is presented and results are found for the scaled down version of the rated load and the light load.
In Chapter 5 the comparison of the mathematical analysis, simulation model and inverter prototype
is shown for the scaled down version which shows agreement among the results found from these
three analyses.
88

6.2 Contributions

The major contributions of the thesis, presented in Chapter 3 and verified through simulation
studies done in Chapters 4 and the prototype study in Chapter 5, are:

1. The loss components are calculated both for maximum rated power and low power
operation using the loss model equations and device datasheet values. Several loss
components (i.e. conduction loss, switching loss, drive loss) that affect inverter efficiency
are taken into account. The switching frequencies are varied from higher to lower values
to see the efficiency trend from rated to low power conditions. The mathematical model
results are then verified by results achieved in simulation and experimental lab
implementation of the proposed inverter.

2. The inverter design is simulated in OrCad PSpice CIS Lite (Version 16.6), to better
understand losses and to accurately predict operation of the inverter in a practical set-up.
The efficiency trends as a function of switching frequency seen in simulations resemble
the efficiency trends in the practical hardware set-up. The simplified load models simulated
in PSpice and the characteristics of the output voltage and current of the simulated inverter
also comply with the practical inverter prototype.

3. Experimentally, a scaled down version of the proposed 200W inverter is implemented in


the lab setup. 12.856W is considered to be the maximum rated power and 4.5% of the rated
89

power is 0.578W is considered as a low power scenario for the lab implementation. The
inverter is tested by changing switching frequencies for both cases with simplified
experimental load models. The current rating is kept the same in the scaled down version
as for the full voltage design, in order to observe the similar inverter efficiency trend as in
the full voltage design (200W rating). A cost effective microcontroller, the MSP3430, is
used to generate the control signals. To suppress the transient voltage spikes at the inverter
output, a snubber circuit is designed. The switching frequencies are kept in a closed range
so that the same snubber circuit can be used for different switching frequencies.

4. The comparison of the efficiency trend results is presented for the loss model based
calculation, computer simulation and the lab based inverter prototype. Issues are discussed
related to the improved inverter efficiency when switching frequencies are decreased at
low power. In mathematical modeling a wide range of switching frequency (200Hz
20kHz) is used to measure the efficiency at rated and low power levels. However, while
measuring efficiency for simulations and hardware implementation the switching
frequency is kept in a more restricted range of 200Hz to 5kHz for the scaled down version
(e.g. 12.856W rating). The restricted range of frequency is employed in order to use the
same snubber circuit for all the switching frequencies.

90

6.3 Suggestions for Future Work

1. The full scale design to provide 50Hz, 220V RMS developed in the mathematical analysis
and for the simulated inverter should be implemented in hardware and a similar analysis
can be performed to verify the efficiency and total loss calculations.

2. To maximize the inverter efficiency at different power levels, there should be an automated
load sensing technique that would sense the load to send feedback to the controller in order
to set the optimum switching frequency.

3. Instead of using the simplified load models in the hardware implantation, real loads (e.g.
CFL, mini refrigerator, sewing machine) can be used to observe the inverter output voltage
and current characteristics.

4. The snubber circuit losses can be included in the loss calculations and the snubber circuits
can be included in the PSpice inverter model. Also, further investigation can be done in
terms of designing an optimum snubber which will perform well through a wide range of
switching frequencies.

5. Further investigation is needed to determine the practical range of switching frequencies.


It is expected that this will not be a trivial task. The higher switching frequencies assist in
reducing current THD for inductive loads but at the cost of reduced inverter efficiency. The

91

lower switching frequencies increase inverter efficiency but with increased current THD
for some loads.

6. This thesis is based on the idea that an inexpensive inverter, for use in the developing
world, may not require an output filter for many household loads. This ideas needs to be
investigated further: loads for which this is true; effect on load performance, load longevity,
etc. Also, if it is found that a filter is necessary, or may be an option for certain loads, an
investigation of filter requirements and means to simplify filter design and reduce filter
cost, would all be helpful.

92

References

[1] J. Muoz and E. Lorenzo, Technical Standard for Stand-Alone PV Systems using Inverters,
in European Photovoltaic Solar Energy Conference, Munich, Germany, 2003, pp. 558-561.
[2] F. Vignola, F. Mavromatakis and J. Krumsick, Performance of PV Inverters, in ASES Annual
Conference, San Diego, CA, 2008.
[3] H. Akca and U. S. Selamogullari, Hybrid Switch use for Light Load Efficiency Improvement
of a Power Inverter for Stand-Alone Residential Renewable Energy System, in European
Conference on Power Electronics and Applications (EPE), IEEE, Lille, France, 2013, pp. 1-6.
[4] B. Johansson, Renewable Energy in Energy-Efficient, Low-Pollution Systems, PhD diss.,
Lund University, 1997.
[5] W. E. Reid, Power Quality Issues- Standards and Guidelines, in IEEE Power Delivery, vol.
32, no. 3, May/June 1996.
[6] IEEE Application Guide for IEEE Std 1547, IEEE Standard for Interconnecting Distributed
Resources with Electric Power Systems, IEEE Standard 1547, Photovoltaics, Dispersed
Generation, 2009.
[7] Characteristics of the Utility Interface for Photovoltaic (PV) Systems, IEC Standard 61727,
International Electro-technical Commission, Report of IEC, 2002.
[8] S. B. Kjaer, J. K. Pedersen and F. Blaabjerg, A Review of Single-Phase Grid-Connected
Inverters for Photovoltaic Modules, in Industry Applications, IEEE, vol. 41, no. 5, pp. 1292-1306,
2005.
[9] Y. Xiong, S. Sun, H. Jia, P. Shea and Z. J. Shen, New Physical Insights on Power MOSFET
Switching Losses, in Power Electronics, IEEE, vol. 24, no. 2, pp. 525-531, 2009.
93

[10] D. Songquan, H. Mao, T. Wu, S. Xiao and I. Batarseh, Power Losses Estimation Platform
for Power Converters, in Applied Power Electronics Conference and Exposition, IEEE, 2004,
Anaheim, CA, pp. 1784-1789.
[11] J. H. Hahn, Modified Sine-Wave Inverter Enhanced, in Power Electronics Technology, vol.
32, no. 8, 2006.
[12] O. Turna, Pure Sine vs. Modified Sine Wave Inverters, San Francisco, CA, April 8, 2011,
[Online],

Available:

http://www.civicsolar.com/resource/pure-sine-vs-modified-sine-wave-

inverters, Access date: September 2013.


[13] F. L. Luo and H. Ye, Advanced DC/AC Inverters: Applications in Renewable Energy, CRC
Press, January 22, 2013.
[14] N. N. Sayeed, A. A. S. Hossain, N. Hasan and M. A. Bhuiyan, Design of a Pure Sine Wave
Inverter for PV Application, PhD diss., BRAC University, 2011.
[15]J. Doucet, D. Eggleston and J. Shaw, DC/AC Pure Sine Wave Inverter, MSc diss.,
NECAMSID, 2007.
[16] P. H. Zope, P. G. Bhangale and S. R. Suralkar, Design and Implementation of Carrier Based
Sinusoidal PWM (Bipolar) Inverter, in International Journal of Engineering Science and
Research, vol. 1, no. 3, December 2012.
[17] D. Homes, T. Lipo, Pulse Width Modulation for Power Converters: Principles and Practice
John Wiley & Sons, Ch. 1, pp. 1-56, 2003.
[18] G. A. Mazur, Digital Multi-meter Principles, in American Technical Publishers, [Online],
Available: http://en-us.fluke.com/training/training-library/measurements/electricity/what-is-dutycycle.html, Access date: November 11, 2014.

94

[19] S. Jian, Dynamics and Control of Switched Electronic Systems, Springer-Verlag London
Limited, Ch. 2, pp. 25-61, 2012.
[20] A. K. Sharma, A. K. Sharma and N. Vijay, Unipolar and Bipolar Voltage Modulation Type
Inverter for Improved Switching Frequencies, in International Journal of Engineering Sciences
and Research Technology, vol. 3, no. 8, August 2014.
[21] D. Li and C. Ramos, Design of a High-Low Unipolar Pulse Width Modulated Inverter,
Senior project, Electrical Engineering Department, California Poly technique State University,
June 2010.
[22] P. N. Swetha and D. S. Reddy, Design and Implementation of Combined Unipolar and
Bipolar PWM Technique for Reducing Current Distortions and Current Ripples. in International
Journal of Computational Science, Mathematics and Engineering, vol. 2, no. 5, May 2015.
[23] R. W. Erickson and D. Maksimovic, Fundamentals of Power Electronics, Springer Science
& Business Media, Ch. 2, pp. 39-61, 2001.
[24] R. Beiranvand, B. Rashidian, M. R. Zolghadri and S. M. H. Alavi, Using LLC Resonant
Converter for Designing Wide-Range Voltage Source, in Industrial Electronics, IEEE
Transactions, vol. 58, no. 5, pp. 1746-1756, 2011.
[25] H. Hu, W. A. Hoor, N. H. Kutkut, I. Batarseh and Z. Shen, Efficiency Improvement of GridTied Inverters at Low Input Power using Pulse-Skipping Control Strategy. in IEEE Transactions
on Power Electronics, vol. 25, no. 12, pp. 3129-3137, December 2010.
[26] V. Yousefzadeh, D. Maksimovi and Q. Lig, A Zero Voltage Switching Single-Phase
Inverter using Hybrid Pulse-Width Modulation Technique, in Power Electronics Specialists
Conference, IEEE, 2004, vol. 2, pp. 1274-1279.

95

[27] B. Wang, X. Xin, S. Wu, H. Wu and J. Ying, Analysis and Implementation of LLC burst
mode for light load efficiency improvement. in Applied Power Electronics Conference and
Exposition, IEEE, Washington, DC, February 2009, pp. 58-64.
[28] E. Seinen, How to Get Higher Efficiency at Lower Loads: A New LLC Platform Makes it
Possible, NXP International Product, March 2015.
[29] S. Maniktala, Understanding and using LLC Converters to Great Advantage, Microsemi,
Analog Mixed Signal Group, 2013.
[30] Z. Zhang, M. Chen, M. Gao, Q. Mo, and Z. Qian, An Optimal Control Method for GridConnected Photovoltaic Micro-inverter to Improve the Efficiency at Light-Load Condition, in
Energy Conversion Congress and Exposition (ECCE), IEEE, Phoenix, AZ, USA, pp. 219-224,
2011.
[31] O. A. Naseem, R. W. Erickson and P. Carlin, Prediction of Switching Loss Variations by
Averaged Switch Modeling, in Applied Power Electronics Conference and Exposition, IEEE,
Dallas, TX, USA, 2000, vol. 1, pp. 242-248.
[32] B. L. C. Martinez, R. Li, K. Ma and D. Xu, Hard Switching and Soft Switching Inverters
Efficiency Evaluation, in International Conference on Electrical Machines and Systems, Wuhan,
China, 2008, pp. 1752-1757.
[33] B. Ismail, S. Taib, M. Isa and C. M. Hadzer, Development of a Single Phase SPWM
Microcontroller-based Inverter, in Power and Energy Conference, IEEE, Gaborone, Botswana,
2006, pp. 437-440.
[34] N. Allet, F. Baumgartner, J. Sutterlueti, S. Sellner and M. Pezzotti, Inverter Performance
under Field Conditions, in EU PVSEC, Hamburg, Germany, 2010, pp. 3110-3117.

96

[35] N. Ekren, N. Onat and S. Saglam, Household Type Loads Effects on Photovoltaic Systems,
in WSEAS Transactions on Circuits and Systems, vol. 7, no. 12, pp. 1020-1028, 2008.
[36] J. H. Lee and B. H. Cho, Large Time-Scale Electro-Thermal Simulation for Loss and
Thermal Management of Power MOSFET, in Power Electronics Specialist Conference, IEEE,
Acapulco, Mexico, June 2003, pp. 112117.
[37] M. C. Oberdorf, Power Losses and Thermal Modeling of a Voltage Source Inverter, PhD
diss., Naval Post Graduate School, Monterey, CA, March 2006.
[38] M .H. Bierhoff and F. W. Fuchs, Semiconductor Losses in Voltage Source and Current
Source IGBT Converters Based on Analytical Derivation, in Power Electronics Specialists
Conference, IEEE, Aachen, Germany, 2004, vol. 4, pp. 2836-2842.
[39] S. K. Pattnaik and K. K. Mahapatra, Power Loss Estimation for PWM and Soft-switching
Inverter using RDCLI, in International Multi Conference of Engineers and Computer Scientists,
Hong Kong, March 17-19, 2010, vol. 2.
[40] C. Graovac and M. Purschel, IGBT Power Losses Calculation using the Data-Sheet
Parameters, Infineon Technologies AG Application Note, January 2009.
[41] K. Berringer, J. Marvin and P. Perruchoud, Semiconductor Power Loss in AC Inverter, in
Industry Applications Conference, Orlando, Florida, USA, 1995, vol. 1, pp. 882-888.
[42] B. K. Bose, Power Electronics and Motor Drives: Advances and Trends Academic press,
Ch. 2, pp. 25-72, 2010.
[43] C. Liu, B. Wu, N. R. Zargari, D. Xu and J. Wang, A Novel Three-Phase Three-Leg AC/AC
Converter using Nine IGBTs, in Power Electronics, IEEE Transactions, vol. 5, no. 24, pp. 11511160, 2009.

97

[44] U. Nicolai and A. Wintrich, Determining Switching Losses of SEMIKRON IGBT Modules,
SEMIKRON Application Note, AN 1403, August 2014.
[45] J. Karlsson, The Concept of IGBT Modeling and the Evaluation of the PSPICE IGBT
Model, MSc. Diss., Industrial and Electrical Engineering and Automation (IEA), 2002.
[46] L. J. Sheng, Resonant Snubber-based Soft-Switching Inverters for Electric Propulsion
Drives, in Industrial Electronics, IEEE Transactions, vol. 44, no. 1, pp. 71-80, 1997.
[47] IGBT/MOSFET Gate Drive Optocoupler, Application Note 91, [Online], Available:
www.vishay.com, access date: November 2014.
[48] N. Chandrasekaran and K. Thyagarajah, Simulation and Experimental Validation of AC
motor and PMDC Motor Pumping System Fed by Photovoltaic Cell, in Indian Journal of
Engineering and Materials Sciences, vol. 21, no. 1, pp. 93-103, February 2014.
[49] S. Maheshri and K. Prabodh, Simulation of Single Phase SPWM Unipolar Inverter, in
International Journal of Innovative Research in Advanced Engineering (IJIRAE), vol. 1, no. 3,
October 2014.
[50] P. H. Zope, P. G. Bhangale, P. Sonare and S. R. Suralkar, Design and Implementation of
Carrier based Sinusoidal PWM Inverter, in International Journal of advanced research in
electrical, electronics and instrumentation engineering, vol. 1, no. 4, pp. 230-236, 2012.
[51] H. L. Chen and C. K. Tee, Keeping Inverters Safe, Avago Technologies San Jose, EET,
[Online], Available: http://powerelectronics.com/power-electronics-systems/keeping-inverterssafe, Access date: September 23, 2014.
[52] B. R. Sidney and D. Holliday, Optimal Regular-Sampled PWM Inverter Control
Techniques, in Industrial Electronics, IEEE Transactions, vol. 54,no. 3, pp. 1547-1559, 2007.

98

[53] S. Jeevananthan, R. Nandhakumar and P. Dananjayan, Inverted Sine Carrier for Fundamental
Fortification in PWM Inverters and FPGA based Implementations, in Serbian Journal of
Electrical Engineering, vol. 4, no. 2, pp. 171-187, 2007.
[54] C. Cresswell, Steady State Load Models for Power System Analysis, Phd diss., May 2009.
[55] W. M. Kwok, Standards of Power Quality with Reference to the Code of Practice for Energy
Efficiency of Electrical Installations, Energy Efficiency Office, Electrical & Mechanical Service
Department, September 2003.
[56] J. Molina and S. Luis, Model of Electronic Ballast Compact Fluorescent Lamps, in Power
Delivery, IEEE, vol. 29, no. 3, pp. 1363-1371, 2014.
[57] N. R. Watson, T. Scott and S. Hirsch, Compact Fluorescent Lamps (CFL) Implications for
Distribution

Networks,

Application

Note

AS/NZS

61000-3-2,

[Online],

Available:

https://www.esmap.org/sites/esmap.org/files/50.%20Neville_Watson_report_on_CFL_Harmonic
s.pdf
[58] K. Deelak, P. Suki and G. tumberger, Impact of Compact Fluorescent Lamps on the
Excessive Transmission Line Losses, in International Conference on Renewable Energies and
Power Quality, Santiago de Compostela, Spain, 2012.
[59] V. George, A. Bagaria, P. Singh, S. R. Pampattiwar and S. Periwal, Comparison of CFL and
LED lamp-harmonic Disturbances, Economics (cost and power quality) and Maximum Possible
loading in a Power System, in Utility Exhibition on Power and Energy Systems: Issues &
Prospects for Asia (ICUE), International Conference, IEEE, Pattaya, Thailand, 2011, pp. 1-5.
[60] Power Quality Implications of Compact Fluorescent Lamps in Residences, National
Electrical

Manufacturers

Association,

99

1999,

[Online],

Available:

http://www.nema.org/Standards/Pages/Power-Quality-Implications-of-Self-ballasted-Lamps-inResidences.aspx, Access date: March 11, 2013.


[61] B. Ledbuizen, Lamp measurement report - CFL E27 30W Germany, in Environmental
Change Institute, January 2013.
[62] A. M. Ahmed, A. A. Bashit, M. S. Alam and H. Begum, Improvement of Input-Side Current
of a Single Phase Rectifier with Variable Output Voltage Range using Boost Converter and
Investigation of Harmonic Minimization, in International Journal of Scientific & Engineering
Research, vol. 4, no. 13, pp. 2229-5518, March 2013.
[63] B. Ireland, The Hidden Costs of CFLs, in Electrical Construction and Maintenance,
[Online], available: http://ecmweb.com/lighting-amp-control/hidden-costs-cfls, Access date:
November 5, 2014.
[64] Code Composer Studio, vol. 6.1, MSP430 User's Guide, may 2009Revised may 2015.
[65] F. Z. Peng, S. J. Gui, L. M. Tolbert, A Passive Soft-Switching Snubber for PWM Inverters,
in Power Electronics, IEEE, vol. 19, no. 2, pp. 363-370, March 2004.
[66] J. Hagerman, Calculating Optimum Snubbers, Hagerman Technology, July 24, 1995.
[67] W. McMurray, Optimum Snubbers for Power Semiconductors, in Industry Applications,
IEEE, pp. 593-600, September 1972.
[68]

RC

Snubbers,

Illinois

Capacitor.

INC,

[Online],

Available:

http://www.illinoiscapacitor.com/pdf/Papers/RC_snubber.pdf, Access date: December 31, 2014.


[69] W. McMurray, Selection of Snubbers and Clamps to Optimize the Design of Transistor
Switching Converters, in Power Electronics Specialists Conference, IEEE, CA, USA, July 1979,
vol. 16, no. 4, pp. 513-523.

100

[70] T. L. Floyd, Electric Circuits Fundamentals, in Prentice Hall, Inc., Ch. 5, pp. 169-188, July
3, 2009.
[71] C. Alexander, C. K. Alexander and M. N. Sadiku, Fundamentals of electric circuits, Urban
Media Comics, pp. 64, 2013.
[72] J. Ware, Power Factor Correction (PFC), IEE Wiring Matters, July 2006.
[73] J. Tsao, N. Lewis, G. Crabtree, Solar FAQs, Sandia National Labs, April 20, 2006.
[74] W. E. Reid, Power Quality Issues- Standards and Guidelines, in IEEE Power Delivery, vol.
32, no. 3, May/June 1996.
[76] J. Dodge and J. Hess IGBT Tutorial, Advanced Power technology, application note
APT0201, EE Times, July 2002.
[77] P. Semig and C. Wells, A Current Sensing Tutorial-Part 1: Fundamentals, Texas
Instruments, EE Times, August 2012.
[78] T. H. Loba, D. H. Wood and E. P. Nowicki, Improving Inverter Efficiency at Low Power by
Reducing Switching Frequency, in Canadian Conference on Electrical and Computer
Engineering (CCECE), IEEE, May 2015, pp. 859-863.

101

APPENDIX A: SCHEMATIC DIAGRAM OF FULL BRIDGE INVERTER

Figure A-1 Full Bridge Inverter Supplying Power to the Heavy Load.

Inverter shown in Figure A-1 is supplying power to the simplified equivalent heavy load model.
L1 and R1 make the equivalent load model. The SPWM controller generates gating signals for the
four switches and controls the switching in such a way that AC signal is delivered at the output of
the inverter. A pair of antiparallel diodes are forward biased based on which switching IGBT pair
is switching.

102

Figure A-3 Full Bridge Inverter Supplying Power to the Light Load.

103

APPENDIX B: COMPONENT LIST FOR EXPERIMENTAL SETUP

The inverter circuit mainly consists of four components:

1. DC-DC converter IC (NME0515SC)


2. Optocoupler IC (FOD3184)
3. IGBT (IRG4BC20UD)
4. Microcontroller development kit (MSP430)

DC-DC Converter IC

NME0515SC is an isolated 1W single output DC/DC converter, shown in Figure B-1. The nominal
input voltage of the converter is 5V and the output voltage is 15V. The maximum output current
is 66mA with conversion efficiency of 80%.

Figure B-1 DC-DC Converter NME0515SC


(Picture from device data sheet: http://datasheet.octopart.com/NME0505SC-Murata-datasheet5308039.pdf)

104

IGBT IC

IRG4BC20UD is an ultra-fast, generation 4 IGBT design co-packaged with anti-parallel diode


shown in Figure B-2. The collector to emitter voltage rating is 600V and the continuous collector
rating is 6.5A. The diode forward voltage drop of the anti-parallel diode is 1.4V for a current of
7A.

Figure B-2 DC-DC Converter NME0515SC


(Picture from device data sheet: http://www.irf.com/productinfo/datasheets/data/irg4bc20ud.pdf)

Optocoupler IC

FOD3184 is a high speed MOSFET/IGBT gate driver optocoupler with 3A peak output current,
shown in Figure B-3. It has a wide output voltage range of 15V to 30V. The input current, , to
turn on the LED is 10mA to 16mA.

105

Figure B-3 Optocoupler FOD3184


(Picture from device data sheet: https://www.fairchildsemi.com/datasheets/FO/FOD3184.pdf)
In this thesis, a MSP430 development kit has been used in order to implement SPWM control. The
details of MSP430 are discussed in Appendix C.

106

APPENDIX C: MICROCONTROLLER (MSP430G2553) OVERVIEW

The microcontroller is the heart of the inverter system. The MSP430G2553 microcontroller IC is
programmed to generate four SPWM signals which are fed to the gates of the IGBTs through the
optocouplers. The chosen microcontroller is small in size, light in weight and inexpensive. Other
considerations for choosing the MSP430G2553 is its high computing power and ultra-low power
consumption.

The MSP430 LaunchPadTM (http of TI website) is an economical microcontroller development


board. The MSP430G2553 IC is embedded in the MSP430 LaunchPad TM (conveniently in an IC
socket). The LaunchPad comes with free software and a full series of tutorials. Also, the MSP430
LaunchPad (Figure A3) includes: a communication module port for computer interfacing, debug
mode for testing the programmes, access to input/output microcontroller's ports, Universal
Asynchronous Receiver/Transmitter Interface (UART) port, and the necessary software and
drivers for the computer.

Figure C-1 MSP430 LaunchPad.


107

The choice of the MSP430G2553 microcontroller is based on: incorporation of an internal


temperature sensor, an integrated Analog-to-Digital Converter (ADC), being programmable in C,
its low price and low power consumption. Its main features are described in the Texas Instruments'
user guide (http://www.ti.com/lit/ug/slau144j/slau144j.pdf). As shown by the arrows in Figure C2, the six pins used in this thesis are: pin 1 , pin 9 P2.1/TA1.1, pin 10 P2.2/TA1.1, pin 11
P2.3/TA1.0, pin 13 P2.4/TA1.2, pin 20 .

Figure C-2 MSP430G2553 Pin Overview

Pin and Timer Description


Referring to Figures C-2 and C-3 a description of relevant pins and timer operation is now
provided.
DVCC: Digital Supply Voltage (3.3V).
P2.1/TA1.1 General - purpose digital input/output pin, Time1_A, capture: CCI1A input, compare:
Out1 output.
P2.2/ TA1.1 General purpose digital input/output pin, Time1_A, capture: CCI1B input, compare:
Out1 output.

108

P2.4/ TA1.2 General purpose digital input/output pin, Time1_A, capture: CC12A input,
compare: Out2 output.
P2.5/TA1.2 General purpose digital input/output pin, Time1_A, capture: CCI2B input, compare:
Out2 output.
DVSS: Digital Ground Reference.
MSP430G2553 has two Timer_A 16 bit timers:
1. Timer0_A
2. Timer1_A
As shown in Figure C-3 different timing modes in MSP430 is shown. Each 16 bit timer starts count
from 0 to 0x0FFFF and they operate in four different modes:
1. Stop: Timer is in halt state or stops the timer.
2. UP: Timer counts up from zero to value stored in TACCR0 register (other than 0xFFFF)
and roll over to zero after it reached the count value as shown in figure C-3. Generally this
mode used to produce time delays.
3. Continuous: It is same as UP mode but here Timer counts up from zero to maximum value
0xFFFFh and rolls over to zero after it reached 0xFFFF and keep going as shown in figure
C-3.
4. Up/Down: In this mode Timer counts up from 0 to TACCR0 register and then counts down
back to zero.

109

Figure C-3 Different timing modes in MSP430 microcontroller.

Figure C-4 Timer register and its configurable bits of MSP430.

110

As shown in Figure C-4, Three compare/capture registers CCR0, CCR1, CCR2 are used to the
timer count. These registers can also be referred as TACCR0, TACCR1, and TACCR2.

1. TAR is the 16-bit timer register in which the count start increment/decrement value
depends upon the timer mode settings.
2. CCIFG interrupt flag is set when the timer counts to the value stored in CCR0
register.
3. TAIFG interrupt flag is set when the timer count from CCR0 to zero.
4. TASSELx are the bits used to select one of the clock signals as shown in the figure
C-4.
5. IDx bits are used to divide the clock signal applied to timer as shown in the figure
C-4.
6. MCx bits are used select count mode as shown in figure C-4.
7. TACLR bit clears the TAR register, clock divider and count direction (mode).

111

APPENDIX D: MICROCONTROLLER CODE

MSP430G2553 is programmed to generate four SPWM signals. To support four SPWM signals,
Timer1_A is used. Corresponding TA1CCR0, TA1CCR1 and TA1CCR2 registers are used for the
carrier-reference compare operation. The code for generating four SPWM signals at 2.5kHz is
provided below as an example.

A SINE lookup table is created using the following parameters:


1. Angle steps - number of slices of full wave period (360 degree);
2. Full range - device full range, for example, 8bit = 256;
3. Waveform Zero - wave value at 0, 180, 360 degree;
4. Half-wave amplitude - wave height at 90 degree;
The sine table can be generated very easily by inserting the above values in
http://www.meraman.com/htmls/en/sinTableOld.html website.

The code has self-explanatory comments following // sign.


// inclusion of microcontroller header file
#include <msp430g2553.h>
// Sinewave frequency is 50Hz which means reference signal period 20ms.
// Triangular wave frequency is 2.5KHz means 0.4ms period. This is also the // period
of a single SPWM pulse.
// That means we have 20/0.4 = 50 SPWM pulses within positive half cycle of
// sine wave.
// Calculation for Setting Up the Compare Registers
// 400 clock ticks with respect to SMCLK (1 MHz) means a period of 0.4ms. CCR0 // is
used for SPWM period. Due to up/down mode CCR0 will be 200.

112

// Variable duty cycle is used for each period of triangle wave to generate
// SPWM signal.
// Hence the sine lookup table will be used to set duty cycle.

// SINE lookup table. Full Range = 200; Angle Steps = 50; Waveform Zero = 100
// Half-Wave Amplitude = 160;
const unsigned int halfcycle[51] = {100, 108, 115, 122, 129, 135, 141, 146, 151, 154,
157, 159, 160, 160, 159, 157, 154, 151, 146, 141, 135, 129, 122, 115, 108, 100, 92,
85, 78, 71, 65, 59, 54, 49, 46, 43, 41, 40, 40, 41, 43, 46, 49, 54, 59, 65, 71, 78,
85, 92 };
// index variable to use for sine lookup table
signed int index = -1;

void main(void)
{
// Stop Watchdog Timer which is not required
WDTCTL = WDTPW + WDTHOLD;
// calibrate Digitally Controlled Oscillator (DCO) for 1MHz operation in
// MSP430G2553
BCSCTL1 = CALBC1_1MHZ;
DCOCTL = CALDCO_1MHZ;
// Setting SPWM PINs
// Port 2 Direction (P2DIR) is set to HIGH for P2.1 (BIT1), P2.2 (BIT2),
// P2.4 (BIT4) and P2.5 (BIT5) to select them as SPWM output
P2DIR |= BIT1 + BIT2 + BIT4 + BIT5;
// Setting pair of active and inactive SPWM PINs. Each pair gets
// activated after Sinewave half-cycle.
P2SEL |= BIT1 + BIT4;
P2SEL &= ~ (BIT2 + BIT5);
// Setting value of SPWM pins when they work together
P2OUT &= ~ (BIT1+BIT2);
P2OUT |= BIT4+BIT5;

// Setting SPMWM period for up/down mode which is 200 with respect to
// 1MHz clock; SPWM period is 0.4ms or 400us

113

TA1CCR0 = 200 - 1;
// CCR0 interrupt. CCR0 interrupt function will be invoked when TA1CCR0
// limit is reached.
TA1CCTL0 = CCIE;
// CCR1 toggle/set. One of the active SPWM signal.
// CCR1 interrupt function will be invoked when CCR1 reaches limit.
TA1CCTL1 = OUTMOD_2 + CCIE;
// CCR2 toggle/reset as complementary to CCR1. The other active SPWM
// signal.
TA1CCTL2 = OUTMOD_6;
// SMCLK (1MHz), up-down mode (MC_3)
TA1CTL = TASSEL_2 + MC_3 + TACLR;
// Enter Low Power Mode
_BIS_SR (LPM0_bits + GIE);
}

// Interrupt Service Routine for CCR1 and CCR2 interrupts


#pragma vector=TIMER1_A1_VECTOR
__interrupt void Timer_A1 (void)
{
switch (__even_in_range(TA1IV, 10))
{
case 0:
// no interrupt is pending
{
break;
}
case 2:
// case 2 means CCR1 interrupt
// The Capture Compare register of Timer A is counting up to 200-1 and then the values
of sine wave table is getting compared with the triangular signal.
{
TA1CCR1 = halfcycle[index] - 1;
TA1CCR2 = halfcycle[index] - 2;
break;

114

}
case 4:
// case 4 means CCR2 interrupt
{
break;
}
case 10:
// case 10 means clearing TA1IFG
{
break;
}
}
}

// Interrupt Service Routine for CCR0


#pragma vector=TIMER1_A0_VECTOR
__interrupt void Timer_A0 (void)
{
// initializing sine look up table index
index = index + 1;
if(index == 25)
{
// Re-setting sine lookup table index to 0 after halfcycle
index = 0;
// Performing XOR to make Active PINs inactive and vice versa
P2SEL ^= (BIT1 + BIT4 + BIT2 + BIT5);
}
}

For other switching frequency values [200Hz and 5kHz in this thesis], the SPWM period changes
and the duty cycle changes as well. By taking appropriate values for the angle steps, full range,
waveform zero and half wave amplitude, the sine table is generated. For 200Hz and 5kHz the code
remains same except for: carrier frequency, SPWM period and the sine table changes. For a
115

switching frequency of 200 Hz: SPWM period, CCR0 = 5-1; for a switching frequency of 5kHz:
CCR0 = 100-1.

116

Você também pode gostar