INVESTIGATION OF HIGH PERFORMANCE SINGLE-PHASE ...

148
INVESTIGATION OF HIGH PERFORMANCE SINGLE-PHASE SOLUTIONS FOR AC-DC POWER FACTOR CORRECTED BOOST CONVERTERS by Fariborz Musavi M.A.Sc., Concordia University, 2001 B.Sc., Iran University of Science and Technology, 1994 A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY in The Faculty of Graduate Studies (Electrical and Computer Engineering) THE UNIVERSITY OF BRITISH COLUMBIA (Vancouver) May 2011 © Fariborz Musavi, 2011

Transcript of INVESTIGATION OF HIGH PERFORMANCE SINGLE-PHASE ...

INVESTIGATION OF HIGH PERFORMANCE SINGLE-PHASE

SOLUTIONS FOR AC-DC POWER FACTOR CORRECTED BOOST

CONVERTERS

by

Fariborz Musavi

M.A.Sc., Concordia University, 2001

B.Sc., Iran University of Science and Technology, 1994

A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF

THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

in

The Faculty of Graduate Studies

(Electrical and Computer Engineering)

THE UNIVERSITY OF BRITISH COLUMBIA

(Vancouver)

May 2011

© Fariborz Musavi, 2011

ii

Abstract

Plug-in Hybrid Electric Vehicles (PHEVs) and Electric Vehicles (EVs) are an emerging

trend in automotive circles, and consumer’s interest is growing rapidly. With the development of

PHEVs, battery chargers for automotive applications are becoming a large market for the power

supply industries. The improvement of overall charger efficiency is critical for the emergence

and acceptance of these vehicular technologies, as the charger efficiency increases, the charge

time and utility cost decreases. Additionally, to meet the efficiency and power factor

requirements and regulatory standards for the AC supply mains, power factor correction is

essential.

Due to limited space in vehicle and increasing power consumption, chargers are required

to deliver more power with smaller volume. As a key component of a charger system, the front-

end AC-DC converter must achieve high efficiency and high power density. In this dissertation,

several conventional plug in hybrid electric vehicle charger front end AC-DC converter

topologies are investigated and a new bridgeless interleaved and a phase shifted semi-bridgeless

power factor corrected converter are proposed to improve the efficiency and performance, which

is critical to minimize the charger size, charging time, and the amount and cost of electricity

drawn from the utility. A detailed analytical model for these topologies is developed, enabling

the calculation of power losses and efficiency. Experimental and simulation results of several

prototype boost converter converting universal AC input voltage to 400 V DC at 3.4 kW are

given to verify the proof of concept, and analytical work reported in this thesis. The results show

a power factor greater than 0.99 from 750 W to 3.4 kW, THD less than 5% from half load to full

load and a peak efficiency of 98.94 % at 265 V input and 1200 W load.

iii

Preface

This thesis contains four chapters that present results that have been published or

submitted for consideration in the form of IEEE refereed papers and scientific journals of which

I am the lead author. The initial project overview was proposed by me. For the scientific journal

submissions and papers, I performed all the research, data analyses, and interpretation of the

results, and prepared the final manuscripts. Co-authors provided advice on methodology and

made editorial comments as required. The complete citations for these papers and the chapters in

which they appear are provided as follows:

Chapter 2 is based on:

[1] F. Musavi ; W. Eberle ; W.G. Dunford, "A High-Performance Single-Phase AC-DC Power

Factor Corrected Boost Converter for plug in Hybrid Electric Vehicle Battery Chargers,"

Published in IEEE Energy Conversion Congress and Exposition, Atlanta, Georgia, 2010.

[2] F. Musavi ; W. Eberle ; W.G. Dunford, "Efficiency Evaluation of Single-Phase Solutions for

AC-DC PFC Boost Converters for Plug-in-Hybrid Electric Vehicle Battery Chargers," Published

in IEEE Vehicle Power and Propulsion Conference, Lille, France, 2010.

[3] F. Musavi ; W. Eberle ; W.G. Dunford, "A High-Performance Single-Phase Bridgeless

Interleaved PFC Converter for Plug-in Hybrid Electric Vehicle Battery Chargers," IEEE

Transactions on Industry Applications, in press, 2010-IPCC-417.R1 (T-IA).

Chapter 3 is based on:

[4] F. Musavi ; W. Eberle ; W.G. Dunford, "A Phase Shifted Semi-Bridgeless Boost Power

Factor Corrected Converter for Plug in Hybrid Electric Vehicle Battery Chargers," Published in

IEEE Applied Power Electronics Conference and Exposition, APEC, Fort Worth, TX, 2011.

iv

Chapter 4 is based on:

[5] F. Musavi ; M. Edington ; W. Eberle ; W.G. Dunford, "Effect of Ripple Steering Technique

on Control Loop Stability of A CCM PFC Boost Converter," Published in IEEE Energy

Conversion Congress and Exposition, Phoenix, AZ, 2011.

Part of Chapter 1, Chapter 2 and Chapter 3 are based on:

[6] F. Musavi ; M. Edington ; W. Eberle ; W.G. Dunford, "Energy Efficiency in Plug in Hybrid

Electric Vehicle Chargers: Evaluation and Comparison of Front End AC-DC Topologies,"

Published in IEEE Energy Conversion Congress and Exposition, Phoenix, AZ, 2011.

[7] D. Gautam ; F. Musavi ; M. Edington ; W. Eberle ; W.G. Dunford, " An Automotive On-

Board 3.3 kW Battery Charger for PHEV Application," Published in IEEE Vehicle Power and

Propulsion Conference, Chicago, IL, 2011.

v

Table of Contents

Abstract ................................................................................................................................................... ii

Preface ................................................................................................................................................... iii

Table of Contents .................................................................................................................................... v

List of Tables ....................................................................................................................................... viii

List of Figures ........................................................................................................................................ ix

List of Symbols .................................................................................................................................... xiv

Acknowledgements ............................................................................................................................. xvii

Dedication .......................................................................................................................................... xviii

CHAPTER 1. Introduction .................................................................................................................. 1

1.1. Introduction ............................................................................................................................. 1

1.2. General Background ................................................................................................................. 3

1.3. Literature Review ..................................................................................................................... 6

1.3.1. Conventional PFC Boost Converter .................................................................................. 8

1.3.2. Bridgeless PFC Boost Converter ..................................................................................... 11

1.3.2.1. Positive "HALF Cycle" Operation .............................................................................. 12

1.3.2.2. Negative "HALF Cycle" Operation ............................................................................. 12

1.3.2.1. PSIM Simulation of Bridgeless Boost Converter ......................................................... 14

1.3.3. Semi-Bridgeless PFC boost Converter ............................................................................ 15

1.3.4. Interleaved PFC Boost Converter .................................................................................... 16

1.3.4.1. PSIM Simulation of Interleaved Boost Converter ........................................................ 18

1.3.5. Ripple Steering Technique in PFC Applications.............................................................. 19

1.3.6. Average Switch Model ................................................................................................... 21

1.4. Motivation and Objectives of the Thesis ................................................................................. 22

1.5. Outline of the Thesis .............................................................................................................. 23

CHAPTER 2. Bridgeless Interleaved PFC Boost Converter ............................................................... 24

2.1. Introduction ........................................................................................................................... 24

2.2. Circuit Description and Steady State Analysis ........................................................................ 24

2.2.1. Positive Half Cycle Operation ........................................................................................ 25

2.2.2. Negative Half Cycle Operation ....................................................................................... 25

2.2.3. Detailed Positive Half Cycle Operation and Analysis for D > 0.5 .................................... 26

vi

2.2.4. Detailed Positive Half Cycle Operation and Analysis for D < 0.5 .................................... 30

2.3. Loss Modeling ....................................................................................................................... 35

2.4. Math Modeling Verification ................................................................................................... 41

2.5. Simulation Results ................................................................................................................. 45

2.6. Logic Implementation ............................................................................................................ 47

2.7. Experimental Results.............................................................................................................. 49

2.8. Conclusion ............................................................................................................................. 57

CHAPTER 3. Phase-Shifted Semi Bridgeless PFC Converter ............................................................ 59

3.1. Introduction ........................................................................................................................... 59

3.2. Circuit Description and Steady State Analysis ........................................................................ 59

3.2.1. Positive Half Cycle Operation ........................................................................................ 60

3.2.2. Negative Half Cycle Operation ....................................................................................... 61

3.2.1. Detailed Positive Half Cycle Operation and Analysis for D > 0.5 .................................... 61

3.2.1. Detailed Positive Half Cycle Operation and Analysis for D < 0.5 .................................... 64

3.3. Loss Modeling ....................................................................................................................... 66

3.4. Simulation Results ................................................................................................................. 68

3.5. Logic Implementation ............................................................................................................ 70

3.6. Experimental Results.............................................................................................................. 71

3.7. Conclusion ............................................................................................................................. 82

CHAPTER 4. The Ripple Steering Technique and Converter Modeling ............................................ 83

4.1. Introduction ........................................................................................................................... 83

4.2. Average Switch Model ........................................................................................................... 83

4.2.1. PWM Switch Model of Conventional Boost Converter ................................................... 83

4.2.1. Feedback Compensation Design of Conventional Boost Converter ................................. 84

4.3. Ripple Steering Techniques in PFC Application ..................................................................... 88

4.3.1. PWM Switch Model of Boost Converter with Coupled Inductor ..................................... 89

4.3.1. Feedback Compensation Design of Boost Converter with Coupled Inductor ................... 92

4.4. Simulation Results ................................................................................................................. 93

4.5. Experimental Results.............................................................................................................. 95

4.6. Conclusion ........................................................................................................................... 102

CHAPTER 5. Conclusions and Future Work ................................................................................... 103

5.1. Conclusions ......................................................................................................................... 103

vii

5.1.1. Bridgeless Interleaved Boost PFC Converter................................................................. 103

5.1.2. Phase Shifted Semi-Bridgeless Boost PFC Converter .................................................... 103

5.1.3. New Loss Modeling for PFC Boost Converters ............................................................. 104

5.1.1. New Average Modeling for PFC Boost Converters with Coupled Inductors .................. 104

5.2. Future Work ......................................................................................................................... 105

5.2.1. LLC Resonant Converter for DC/DC Stage................................................................... 105

5.2.2. Resonant PFC Converter .............................................................................................. 105

5.2.3. Level 3 Chargers .......................................................................................................... 105

5.2.4. Wireless Chargers......................................................................................................... 106

Bibliography ....................................................................................................................................... 107

Appendix ............................................................................................................................................ 112

viii

List of Tables

Table 2-1: Summary of component RMS current for conventional boost and bridgeless boost topologies

...................................................................................................................................................... 37

Table 2-2: Summary of component RMS current for interleaved and bridgeless interleaved boost

topologies ...................................................................................................................................... 37

Table 2-3: Summary of component average current for conventional boost and bridgeless boost

topologies ...................................................................................................................................... 38

Table 2-4: Summary of component average current for interleaved and bridgeless interleaved boost

topologies ...................................................................................................................................... 38

Table 2-5: Component / devices used in prototype unit .......................................................................... 52

Table 3-1: Component / devices used in prototype unit .......................................................................... 71

ix

List of Figures

Figure 1-1: Simplified system block diagram of a universal two-stage battery charger ............................. 3

Figure 1-2: Passive power factor correction AC main voltage and current waveforms .............................. 5

Figure 1-3: Conventional PFC boost converter......................................................................................... 8

Figure 1-4: Simplified block diagram of the boost PFC circuit ................................................................. 9

Figure 1-5: Inductor current and the duty cycle in a typical PFC boost converter in CCM ...................... 10

Figure 1-6: Top: Transistor current - Bottom: Diode current in a typical PFC boost converter in CCM ... 10

Figure 1-7: Bridgeless PFC boost topology ............................................................................................ 11

Figure 1-8: Bridgeless PFC converter operation ..................................................................................... 13

Figure 1-9: PSIM simulation circuit for bridgeless PFC boost converter: Vin = 240V, Vo = 400V, Po =

3400W and fsw = 70 kHz .............................................................................................................. 14

Figure 1-10: Simulation waveforms for bridgeless PFC boost converter including output voltage, input

voltage and input current: Vin = 240V, Vo = 400V, Po = 3400W and fsw = 70 kHz ....................... 15

Figure 1-11: Bridgeless PFC boost topology .......................................................................................... 16

Figure 1-12: Interleaved PFC boost topology ......................................................................................... 17

Figure 1-13: PSIM simulation circuit for interleaved PFC boost converter: Vin = 240V, Vo = 400V, Po =

3400W and fsw = 70 kHz .............................................................................................................. 18

Figure 1-14: Simulation waveforms for interleaved PFC boost converter including output voltage, input

voltage and input current: Vin = 240V, Vo = 400V, Po = 3400W and fsw = 70 kHz ....................... 19

Figure 1-15: Smoothing transformer in a coupled filter .......................................................................... 20

Figure 1-16: Modified PFC boost converter with coupled inductors ....................................................... 20

Figure 1-17: One active switch and one passive switch .......................................................................... 21

Figure 1-18: PWM-switch ..................................................................................................................... 21

Figure 1-19: Averaged model of PWM-switch ....................................................................................... 21

Figure 2-1: Proposed bridgeless interleaved (BLIL) PFC boost converter ............................................... 25

Figure 2-2: Interval 1: Q1 and Q2 are “ON”, and body diode of Q4 conducting ..................................... 26

Figure 2-3: Interval 2 and 4: Q1, Q2, Q3 and Q4 are “ON” .................................................................... 26

Figure 2-4: Interval 3: Q3 and Q4 are “ON”, and body diode of Q2 conducting ..................................... 27

Figure 2-5: BLIL PFC boost converter steady-state waveforms at D > 0.5.............................................. 28

Figure 2-6: Interval 1 and 3: Body diodes of Q2 and Q4 conducting ...................................................... 31

Figure 2-7: Interval 2: Q1 and Q2 are “ON”, and body diode of Q4 conducting ..................................... 31

Figure 2-8: Interval 4: Q3 and Q4 are “ON”, and body diode of Q2 conducting ..................................... 31

Figure 2-9: BLIL PFC boost converter steady-state waveforms at D < 0.5.............................................. 33

x

Figure 2-10: RMS ripple current through output capacitors vs input voltage at Po = 3400 W, fsw = 70

kHz ............................................................................................................................................... 39

Figure 2-11: Comparison of the estimated loss distribution in the semiconductors at 70kHz switching

frequency, 240V input, 3400W load at 400V ................................................................................. 40

Figure 2-12: Simulation and math modeling of inductor RMS current for an interleaved PFC converter . 42

Figure 2-13: Simulation and math modeling of MOSFET RMS current for an interleaved PFC converter

...................................................................................................................................................... 42

Figure 2-14: Simulation and math modeling of boost diode average current for an interleaved PFC

converter ....................................................................................................................................... 43

Figure 2-15: Simulation and math modeling of inductor RMS current for a bridgeless PFC converter .... 43

Figure 2-16: Simulation and math modeling of MOSFET RMS current for a bridgeless PFC converter .. 44

Figure 2-17: Simulation and math modeling of boost diode average current for a bridgeless PFC converter

...................................................................................................................................................... 44

Figure 2-18: Simulation and math modeling of MOSFET body diode average current for a bridgeless PFC

...................................................................................................................................................... 45

Figure 2-19: PSIM simulation circuit for the proposed BLIL PFC boost converter: Vin = 240V, Vo =

400V, Po = 3400W and fsw = 70 kHz ............................................................................................ 46

Figure 2-20: Simulation waveforms for the proposed BLIL PFC boost converter including output voltage,

input voltage and input current: Vin = 240V, Vo = 400V, Po = 3400W and fsw = 70 kHz .............. 47

Figure 2-21: Logic implementation of BLIL PFC boost converter using UCC28070 controller .............. 48

Figure 2-22: Breadboard prototype of BLIL PFC boost converter .......................................................... 49

Figure 2-23: Breadboard proposed BLIL PFC experimental waveforms; ................................................ 50

Figure 2-24: Gating signal, boost MOSFET and diode current for D > 0.5 ............................................. 50

Figure 2-25: Gating signal, boost MOSFET and diode current for D < 0.5 ............................................. 51

Figure 2-26: Load transient response from FL to NL (3400W to 0W) .................................................... 51

Figure 2-27: Load transient response from NL to FL (0W to 3400 W).................................................... 52

Figure 2-28: Input current harmonics at full load for Vin = 120 V at Po = 1700 W and 240 V at Po = 3400

W .................................................................................................................................................. 53

Figure 2-29: RMS Total harmonics distortion vs. output power at Vin = 120V and Vin = 240V ............. 53

Figure 2-30: Power factor vs. output power at Vin = 120V and Vin = 240V ........................................... 54

Figure 2-31: Efficiency vs. output power for the proposed bridgeless interleaved PFC boost converter .. 54

Figure 2-32: Efficiency vs. output power for different topologies at Vin = 240V .................................... 55

Figure 2-33: Loss reduction as a function of output power at Vin = 240V, Vo = 400 V and 70 kHz for the

proposed BLIL converter compared to the benchmark interleaved boost converter ......................... 56

xi

Figure 2-34: Loss reduction as a function of output power at Vin = 120V, Vo = 400 V and 70 kHz for the

proposed BLIL converter compared to the benchmark interleaved boost converter ......................... 57

Figure 3-1: Phase shifted semi-bridgeless PFC boost topology ............................................................... 60

Figure 3-2: Interval 1and 3: Q1 and Q2 are ON...................................................................................... 61

Figure 3-3: Interval 2: Q1 ON, body diode of Q2 conducting ................................................................. 62

Figure 3-4: Interval 4: Q1 OFF and Q2 ON............................................................................................ 62

Figure 3-5: Phase shifted semi-bridgeless boost converter steady-state waveforms at D > 0.5................. 63

Figure 3-6: Interval 1and 3: Q1 and Q2 are OFF, body diode of Q2 conducting ..................................... 64

Figure 3-7: Interval 2: Q1 ON, body diode of Q2 conducting ................................................................. 64

Figure 3-8: Interval 4: Q1 OFF and Q2 ON............................................................................................ 64

Figure 3-9: Phase shifted semi-bridgeless boost converter steady-state waveforms at D < 0.5................. 65

Figure 3-10: Comparison of the estimated loss distribution in the semiconductors at 70kHz switching

frequency, 240V input, 3300W load at 400V ................................................................................. 67

Figure 3-11: PSIM simulation circuit for the phase shifted semi-bridgeless PFC boost converter: Vin =

240V, Vo = 400V, Po = 3400W and fsw = 70 kHz ......................................................................... 68

Figure 3-12: Simulation waveforms for the proposed phase shifted semi-bridgeless PFC boost converter:

Output voltage, input voltage and input current: Vin = 240V, Vo = 400V, Po = 3400W and fsw = 70

kHz ............................................................................................................................................... 69

Figure 3-13: Logic implementation of phase shifted semi-bridgeless PFC boost converter using

UCC28070 controller ..................................................................................................................... 70

Figure 3-14: Capacitor bank of 820 uF................................................................................................... 72

Figure 3-15: Control board of PFC boost converter ................................................................................ 72

Figure 3-16: IMS power board attached to a heatsink with the PFC inductors ........................................ 73

Figure 3-17: Efficiency as a function of output power at Vin = 240V, Vo=400V and 70kHz switching .. 74

Figure 3-18: Loss reduction as a function of output power at Vin = 240V, Vo=400V and 70kHz for the

proposed phase shifted semi-bridgeless converter compared to the benchmark interleaved boost

converter ....................................................................................................................................... 74

Figure 3-19: Efficiency as a function of output power at Vin = 120V, Vo=400V and 70kHz .................. 75

Figure 3-20: Loss reduction as a function of output power at Vin = 120V, Vo=400V and 70kHz for the

proposed phase shifted semi-bridgeless converter compared to the benchmark interleaved boost

converter ....................................................................................................................................... 75

Figure 3-21: THD as a function of output power at Vin = 120 V and 240V, Vo=400V and 70kHz ......... 76

Figure 3-22: Power factor as a function of output power at Vin = 120 V and 240V, Vo=400V and 70kHz

...................................................................................................................................................... 76

xii

Figure 3-23: Harmonics orders at Vin = 120 V and 240V, compared against EN61000-3-2 standard for

Vin = 120 V at Po = 1700 W and 240 V at Po = 3400 W ................................................................ 77

Figure 3-24: Input current, input voltage and output voltage. ................................................................. 78

Figure 3-25: Input current, inductor current, input voltage and output voltage. ....................................... 79

Figure 3-26: Inductor current, input current and sensed MOSFET current. ............................................. 79

Figure 3-27: Gating signal, inductor and sensed MOSFET current for D < 0.5 ....................................... 80

Figure 3-28: Gating signal, inductor and sensed MOSFET current for D > 0.5 ....................................... 80

Figure 3-29: Load transient response from NL to FL (0W to 3400 W).................................................... 81

Figure 3-30: Load transient response from FL to NL (3400 W to 0 W)................................................... 81

Figure 4-1: Conventional boost converter with PWM switch.................................................................. 84

Figure 4-2: Current loop plant and compensator type II Bode plots ........................................................ 85

Figure 4-3: Type II compensator network .............................................................................................. 85

Figure 4-4: Open loop Bode plot for current loop ................................................................................... 86

Figure 4-5: Voltage loop plant and compensator type II Bode plots ........................................................ 87

Figure 4-6: Open loop Bode plot for voltage loop .................................................................................. 88

Figure 4-7: Modified PFC boost converter with coupled inductors ......................................................... 88

Figure 4-8: Equivalent circuit of coupled inductors ................................................................................ 89

Figure 4-9: Equivalent circuit of coupled inductors ................................................................................ 89

Figure 4-10: Modified boost converter with PWM switch ...................................................................... 90

Figure 4-11: Current loop plant Bode plots ............................................................................................ 91

Figure 4-12: Current loop plant and compensator Bode plots for boos converter with coupled inductor .. 92

Figure 4-13: Open loop plant Bode plot for boost converter with coupled inductor ................................. 93

Figure 4-14: PSIM simulation circuit for ripple steering technique applied to PFC boost converter ........ 94

Figure 4-15: Inductor Current - no filtering technique (Top) – ripple steering technique (Middle), series

capacitor current (Bottom): Vin = 240 V, Vo = 400 V, Po = 1600 W, fsw = 70 kHz ....................... 94

Figure 4-16: Coupled inductors used in experimental circuit - Top left: DC inductor, Top right: AC

inductor, Bottom: Coupled inductor ............................................................................................... 95

Figure 4-17: Inductor current Idc ripple at 120 V input and 800 W output - Experimental ...................... 96

Figure 4-18: Inductor current Idc ripple at 120 V input and 800 W output - Simulation .......................... 96

Figure 4-19: Inductor current Iac at 120 V input and 800 W output - Experimental ................................ 97

Figure 4-20: Inductor current Iac at 120 V input and 800 W output - Simulation .................................... 97

Figure 4-21: Peak inductor current Iac ripple at 120 V input and 800 W output - Experimental .............. 98

Figure 4-22: Peak inductor current Iac ripple at 120 V input and 800 W output - Simulation .................. 98

Figure 4-23: Inductor current Idc ripple at 240 V input and 1600 W output - Experimental .................... 99

xiii

Figure 4-24: Inductor current Idc ripple at 240 V input and 1600 W output - Simulation ........................ 99

Figure 4-25: Inductor current Iac at 240 V input and 1600 W output - Experimental ............................ 100

Figure 4-26: Inductor current Iac at 240 V input and 1600 W output - Simulation ................................ 100

Figure 4-27: Peak inductor current Iac ripple at 240 V input and 1600 W output - Experimental .......... 101

Figure 4-28: Peak inductor current Iac ripple at 240 V input and 1600 W output - Simulation .............. 101

xiv

List of Symbols

CO Output Capacitor

dB/dec Decibel per Decade

δD Boost Diode Duty Cycle

δQ Boost Main Switch Duty Cycle

Irms Total Line Current

I1rms Fundamental Component

Kd Distortion Factor

Kθ Displacement Factor

Rs Current Sense Resistor

TS One Switching Period

θ1 Angle between Current Fundamental and Sinusoidal Line Voltage

VS Maximum Input Voltage

Vref Reference Voltage

Abbreviations

AC Alternating Current

ADC Analog to Digital Converter

ASIC Application Specific Integrated Circuits

BJT Bipolar Junction Transistor

BL Bridgeless

BLIL Bridgeless Interleaved

CCM Continuous Conduction Mode

CRM Critical Conduction Mode

D Duty Cycle

xv

DC Direct Current

DCM Discontinuous Conduction Mode

DSP Digital Signal Processors

FL Full Load

FPGA Field Programmable Gate Array

IC Integrated Circuits

IGBT Insulated Gate Bipolar Transistor

IL Interleaved

IMS Insulated Metal Substrate

EMI Electrical Magnetic Interference

EV Electric Vehicle

HS High Side

LLC Inductor Inductor Capacitor

M Mutual Inductance

MOSFET Metal Oxide Silicon Field Effect Transistor

NL No Load

PID Proportional Integral Derivative

PCB Printed Circuit Board

PF Power Factor

PFC Power Factor Correction

PHEV Plug in Hybrid Electric Vehicle

PSFB Phase Shift Full-Bridge

PWM Pulse Width Modulation

RLC Resistance Inductor Capacitor

RMS Root Mean Square

SR Synchronous Rectifier

xvi

THD Total Harmonics Distortion

ZCS Zero Current Switching

ZVS Zero Voltage Switching

Prefixes for SI Units

p Pico (10-12

)

n Nano (10-9

)

µ Micro (10-6

)

m Milli (10-3

)

k Kilo (103)

M Mega (106)

G Giga (109)

SI Units

A Amperes

C Coulombs

F Farads

H Henries

Hz Hertz

s seconds

V Volts

W Watts

° Degrees

Ω Ohms

xvii

Acknowledgements

I would like to thank my supervisors, Dr. William G. Dunford and Dr. Wilson Eberle for

their guidance, encouragement and continuous support through the course of this work. Their

knowledge, research attitude and ways of thinking are greatly appreciated.

Financial and technical support in the form of project funding, lab equipment and tuition

fee reimbursement from Delta-q Technologies Corp. is greatly acknowledged and appreciated.

I would like to thank all of my past and present colleagues from Delta-q Technologies.

In particular, I would like to thank Ken Fielding, President & CEO, Rob Cameron, former CTO

and VP of Engineering, Art Gau, Senior Design Engineer and Deepak Gautam Power Electronics

Engineer at Delta-Q Technologies Corp. for all their support and valuable discussions on several

topics.

It has been a great pleasure to work in the UBC Electric Power and Energy Systems

Group. I would like to acknowledge the group administrative and management staff.

I would like to give my special thanks to all my family; their love and support make my

life more colorful and meaningful.

xviii

Dedication

To my family

1

CHAPTER 1. Introduction

1.1. Introduction

As the demand for energy drastically increased in the 20th

Century, fossil fuels became

the main source of energy due to convenience and cost. Over the years, however, the price of oil

and problems caused by pollution, have increased considerably, putting pressure on governments

and industries to invest on other solutions to replace fossil fuels. Consequently, interest in other

means of transportation, such as Plug in Hybrid Electric Vehicles (PHEV) and Electric Vehicles

(EV), has increased again.

EV and PHEV technology has existed since the early 1900s. However, the high cost and

low energy density of available energy storage systems, primarily batteries, along with the very

low cost of oil, had limited the interest in EV and PHEV. Recent innovations in lithium-ion

batteries, the higher price of gas, and the air pollution associated with fossil fuels have

significantly impacted the alternative transportation industry.

As the adoption rate of these vehicles increases, the stress on the utility grid is projected

to increase significantly at times of peak demand [1]. Therefore, efficient and high power factor

charging is critical in order to minimize the utility load stress, and reduce the charging time. In

addition, a high power factor is needed to limit the input current harmonics drawn by these

chargers and to meet regulatory standards, such as IEC 61000-3-2 [2].

A PHEV is a hybrid vehicle with a storage system that can be recharged by connecting a

plug to an external electric power source. The charging AC outlet inevitably needs an on-board

AC/DC charger with a power factor correction. An on-board 3.5 kW charger could charge a

depleted battery pack in PHEVs to 95% charge in about four hours from a 240 V supply [3].

2

Chargers are also classified by the level of power they can provide to the battery pack [4]:

• Level 1: Common household circuit, rated to 120 volts AC and 15 amperes. These

chargers use the standard three-prong household connection, and they are usually

considered portable equipment.

• Level 2: Permanently wired electric vehicle supply equipment used especially for

electric vehicle charging; rated up to 240 volts AC, up to 60 amps, and up to 14.4

kilowatts.

• Level 3: Permanently wired electric vehicle supply equipment used especially for

electric vehicle charging; rated greater than 14.4 kW. Fast chargers are rated as Level

3, but not all Level 3 chargers are fast chargers. This designation depends on the size

of the battery pack to be charged and how much time is required to charge the battery

pack. A charger can be considered a fast charger if it can charge an average electric

vehicle battery pack in 30 minutes or less.

The front-end AC-DC converter is a key component of the charger system. The purpose

of this document is to illustrate how this research will be conducted on the high-performance

single-phase solutions for AC-DC power factor corrected converters for PHEV battery chargers.

A variety of circuit topologies and control methods have been developed for the PFC

application [5-7]. The single-phase active PFC techniques can be divided into two categories: the

single-stage approach and the two-stage approach. The single-stage approach is suitable for low

power applications. In addition, due to large low frequency ripple in the output current, only lead

acid batteries are chargeable. Therefore, the two-stage approach is the proper candidate for

PHEV battery chargers [8], where the power rating is relatively high, and lithium-ion batteries

3

are used as the main energy storage system. The front end PFC section is then followed by a

DC/DC section to complete the charger system.

Figure 1-1 illustrates a simplified block diagram of a universal input two-stage battery

charger used for PHEVs.

Figure 1-1: Simplified system block diagram of a universal two-stage battery charger

The PFC stage rectifies the input AC voltage and transfers it into a regulated intermediate

DC link bus. At the same time, power factor correction function is achieved. The following

DC/DC stage then converts the DC bus voltage into a regulated output DC voltage for charging

batteries, which is required to meet the regulation and transient requirements.

1.2. General Background

According to the requirements of input current harmonics [9] and output voltage regulation, a

front-end converter is normally implemented by a power factor correction (PFC) stage.

Conventionally, most of the power conversion equipment employs either a diode rectifier or a

thyristor rectifier with a bulk capacitor to converter AC voltage to DC voltage before processing

it [10]. Such rectifiers produce input current with rich harmonic content, which pollute the power

system and the utility lines. Power quality is becoming a major concern for many electrical users.

The simplest form of PFC is passive (Passive PFC). A passive PFC uses a filter at the AC input

to correct poor power factor. The passive PFC circuitry uses only passive components - an

4

inductor and some capacitors. Although pleasantly simple and robust, a passive PFC rarely

achieves low Total Harmonic Distortion (THD). Also, because the circuit operates at the low line

power frequency of 50Hz or 60Hz, the passive elements are normally bulky and heavy. Figure 1-

2 shows input voltage and current for a passive PFC, and the harmonic spectrum of input current.

The input power factor (PF) is defined as the ratio of the real power over apparent power as:

1-1

Assuming an ideal sinusoidal input voltage source, the power factor can be expressed as the

product of two factors, the distortion factor and the displacement factor, as given:

1-2

The distortion factor Kd is the ratio of the fundamental RMS current to the total RMS current.

a) Input voltage and Input current

Input Voltage

Input Current

5

b) Harmonic Spectrum of Input Current

Figure 1-2: Passive power factor correction AC main voltage and current waveforms

The displacement factor, Kθ, is the cosine of the displacement angle between the fundamental

input current and the input voltage fundamental RMS current.

!"# !"# 1-3

$%& 1-4

where I1rms is the fundamental component of the line current, Irms is the total line current, and θ1

is the phase shift of the current fundamental relative to the sinusoidal line voltage.

The distortion factor is close to unity, even for waveforms with noticeable distortion; therefore, it

is not a very convenient measure of distortion for practical use. The distortion factor is uniquely

related to another figure of merit; the total harmonic distortion (THD):

'() * !"#+, !"#+ !"#+ 1-5

* &&-./0+ 1-6

3rd

Harmonic

1st

Fundamental

5th

Harmonic

6

Kd is regulated by IEC 1000-3-2 [2] for lower power levels and by IEEE Std 519-1992 [11] for

higher power levels, where Kθ is regulated by utility companies [12].

Significant reduction of current harmonics in single-phase circuits can only be achieved

by using rectifiers based on pulse width modulated (PWM) switching converters. These

converters can be designed to emulate a resistive load and, therefore, produce very little

distortion of the current. By using PWM or other modulation techniques, these converters draw a

nearly sinusoidal current from the ac line in phase with the line voltage. As a result, the rectifier

operates with very low current harmonic distortion and very high, practically unity power factor.

This technique is commonly known as power factor correction (PFC). As a result of this

research, the existing PFC technology based on the boost converter topology with average-

current-mode control was significantly improved. The proposed improvements allowed an

extended range of operating conditions and additional functionality.

1.3. Literature Review

All basic power converter topologies, such as boost, buck [13], buck-boost, and their

variations, can be used to realize active PFC techniques [14, 15]. At lower power ratings,

MOSFETs [16] are the switching power devices of choice because of their low conduction losses

and high switching speed. For medium and high-power applications, IGBTs can be used in

PWM-controlled converters with switching frequency of up to 30 kHz. There are many

integrated circuits (ICs) [17] on the market that incorporate control functions for PFC converters

and facilitate compact and cost-effective designs. Digital signal processors (DSPs) [18-21] and

microcontrollers [22] have been successfully used to control PFC converters. Microcontrollers

and DSPs can be used to realize traditional [23, 24] proportional-integral derivative (PID) control

laws as well as non-traditional control principles, such as sliding mode control [25], fuzzy logic

7

[26], and neural networks [27]. One of the drawbacks of using microcontrollers and DSPs is the

significant effort that goes into software development.

The boost topology is by far the most popular topology in PFC applications. The boost PFC

converter draws a continuous current from the line and, therefore, does not require much

filtering, which is usually accomplished by an input filter capacitor. Other topologies such as

buck, buck-boost, and flyback draw pulsed current and need a much better input filter. Unlike the

buck topology, the boost converter easily accommodates the input voltage range, from zero to

the line peak voltage.

The boost converter can operate in continuous conduction mode (CCM) [28], discontinuous

conduction mode (DCM), or critical conduction mode (CRM). These names refer to the

continuity of the inductor current within the switching cycle [29]. The boost converter operating

in DCM and CRM modes is usually easier to control, but it has higher peak-to-peak current

ripple, which causes higher RMS value of the inductor current, higher magnetic and conduction

losses, and higher switching noise, which leads to increased filtering requirements. Therefore,

these modes are restricted to relatively low power levels, while the CCM is used at medium and

high power levels.

While the discontinuous conduction mode (DCM) converters such as boost and flyback

converters are well suited for low power applications, continuous conduction mode (CCM) boost

converters with average current mode [30], peak current mode [31] or hysteresis control [32] are

commonly chosen for many medium and high power applications. The output voltage of the

boost PFC converter should be always higher than the peak line voltage. For universal line

application (85 V-265 V), the output voltage is usually set around 400 VDC [33].

8

The boost circuit-based PFC topology operated in CCM is employed in this study as the main

candidate for front end single-phase solutions for AC-DC power factor corrected converters used

in PHEV battery chargers.

1.3.1. Conventional PFC Boost Converter

The conventional boost topology is the most popular topology for PFC applications. It

uses a dedicated diode bridge to rectify the AC input voltage to DC, which is then followed by

the boost section, as shown in Figure 1-3.

Figure 1-3: Conventional PFC boost converter

The simplified block diagram of the boost PFC circuit is shown in Figure 1-4 [34]. This circuit

has two control loops: One is the fast acting internal current loop. It defines the input current

shape to be sinusoidal and forces it in phase with the input voltage. The other is the external slow

voltage loop which regulates the output dc voltage. The voltage loop should not react to the

120Hz rectified mains variations, so its bandwidth is between 10 to 20 Hz. The current loop

usually has a bandwidth frequency of less than one tenth of the switching frequency.

9

D1

D3 D4

D2

QB

LB DB

Co

Vin

RL

Rs

Rvi

Rvd

PWM

Drive

Multiplier

Power StageAC Source & RectifierOutput Filter, Load &

Voltage Divider

Full wave

Rectified AC

Input Sensing

Current

Sensing Output

Voltage

SensingVoltage

Error

Vref

Figure 1-4: Simplified block diagram of the boost PFC circuit

The principle operation of boost PFC is as follows: The rectified sinusoidal input voltage

goes to a multiplier circuit, providing a current reference to the multiplier and a feedforward

signal proportional to the RMS value of the line voltage. The filtered dc output voltage of the

boost PFC is compared to a reference voltage, Vref and amplified. The error amplifier senses the

variations between the output voltage and the fixed dc reference voltage. The error signal is then

applied to the multiplier. The multiplier's output follows the shape of the input ac voltage, with

an average value inversely proportional to the RMS value of the ac input voltage. This signal is

compared to the current signal sensed by Rs in a pulse-width modulation (PWM) circuit. The

inductor current waveform follows the shape of the rectified ac line voltage. The gate drive

signal controls the inductor current amplitude and maintains a constant output voltage.

As can be seen in Figure 1-5 and Figure 1-6, the inductor current follows the shape of the

input voltage. The transistor and diode current waveforms in a typical boost converter in

10

continuous conduction mode of operation are pulsed-width modulated, with both the duty cycle

and the peak amplitude varying with the ac input voltage.

Figure 1-5: Inductor current and the duty cycle in a typical PFC boost converter in CCM

Figure 1-6: Top: Transistor current - Bottom: Diode current in a typical PFC boost converter in CCM

The inductor ripple current is directly seen at the converter’s input and will require

filtering to meet EMI specifications [35-46]. The diode output current is discontinuous and needs

Boost Converter

Duty Cycle

Boost Converter

Inductor Current

Boost Transistor

Current

Boost Diode

Current

11

to be filtered out by the output capacitor Co. In this topology, the output capacitor ripple current

is very high and its value is the difference between diode current and the dc output current [47].

In practical applications as the power level increases, the diode bridge losses become

significant, so dealing with heat dissipation in a limited surface area is important, particularly

from an efficiency point of view. Therefore, the conventional PFC boost is limited to a low to a

medium power range (e.g. less than 1000 W).

1.3.2. Bridgeless PFC Boost Converter

The bridgeless PFC (BL PFC) boost converter avoids the need for the rectifier input bridge

yet maintains the classic boost topology, as shown in Figure 1-7 [48-57]. As it is demonstrated in

the following, BL PFC does not have any advantages over conventional PFC in terms of passive

components sizing. However, it eliminates the input bridge rectifiers, so it is an attractive

solution for higher power applications, where the dissipating power in a small area is becoming

extremely difficult and overall efficiency is very important.

Figure 1-7: Bridgeless PFC boost topology

The intrinsic body diode connected between the drain and source of the Power MOSFET

switches has an important role in this topology. The circuit shown from a functional point of

view is similar to the common boost converter. In the traditional topology current flows through

12

two of the bridge diodes in series. In the bridgeless PFC configuration, current flows through

only one diode with the Power MOSFET providing the return path. To analyze the circuit

operation, it is necessary to separate it into two sections. The first section operates as the boost

stage and the second section operates as the return path for the AC input signal.

1.3.2.1. Positive "HALF Cycle" Operation

When the AC input voltage goes positive, the gate of Q1 is driven high and current flows

from the input through the inductors, storing energy in L1 and L2 as shown in Figure 1-8-a).

When Q1 turns off, energy in the inductors is released as current flows through D1, through the

load and returns through the body diode of Q2 back to the input mains. During the-off time, the

current through the inductors (which during this time discharges its energy) flows into the boost

diode D1 and close the circuit through the load.

1.3.2.2. Negative "HALF Cycle" Operation

During the negative half cycle circuit operation is mirrored. Q2 turns on, current flows

through the inductor, storing energy, as shown in Figure 1-8-b). When Q2 turns off, energy is

released as current flows through D2, through the load, and back to the mains through the body

diode of Q1.

a) Positive half cycle

13

b) Negative half cycle

Figure 1-8: Bridgeless PFC converter operation

Note that the two Power MOSFETs are driven synchronously. It doesn't matter whether

the sections are performing as an active boost or as a path for the current to return. In either case

there is a benefit of lower power dissipation when current flows through the Power MOSFETs

(through intrinsic body diodes) during the return phase.

As can be noted, in bridgeless topology a new loss has been introduced in the intrinsic

body diodes of MOSFETs and it might add extra power dissipation in the MOSFETs, but since

input bridge rectifiers were eliminated, there is some efficiency gain in overall performance of

the bridgeless topology.

For a conventional PFC, the current sense is easy to monitor by simply inserting a shunt

sensing resistor at the return path of the inductor current. However, for a bridgeless PFC, current

path does not share the same ground at each half-line cycle. A sensing-power MOSFET and

diode current are needed, which makes the bridgeless PFC’s current sensing complicated and

difficult to monitor [52, 58].

In addition, since the AC line is floating compared to the PFC stage ground, simple

circuitry cannot sense input voltage. Normally a low-frequency transformer or optical coupler is

required to perform input voltage sensing.

14

1.3.2.1. PSIM Simulation of Bridgeless Boost Converter

PSIM simulation software was used to verify the steady state waveforms of each component.

Figure 1-9 shows the PSIM simulation circuits of the bridgeless boost PFC converter. As it can

be seen, the power stage section of converter consists of two boost inductors, Ld1 and Ld2, two

fast boost diodes, Db1 and Db2, two MOSFETs, Q1 and Q2 and their body diodes. Also it

consists of one current loop and one voltage loop. The sensed input voltage is multiplied by the

compensated output voltage, and then generates a control signal to be compared with the

switching carrier waveforms to generate the gating signals for main MOSFETs.

Figure 1-9: PSIM simulation circuit for bridgeless PFC boost converter: Vin = 240V, Vo = 400V, Po = 3400W

and fsw = 70 kHz

15

Figure 1-10 shows the PSIM simulation results of a bridgeless PFC boost converter. The

input current is in phase with input voltage, and it has close to unity power factor. Also the

output voltage is regulated at around 400V, with a 120 Hz low frequency ripple. The converter is

operating at 70 kHz switching frequency, 240 V input voltage and 3.4 kW output power.

Figure 1-10: Simulation waveforms for bridgeless PFC boost converter including output voltage, input

voltage and input current: Vin = 240V, Vo = 400V, Po = 3400W and fsw = 70 kHz

1.3.3. Semi-Bridgeless PFC boost Converter

The main disadvantage of bridgeless boost converter is the EMI noise issues [54, 59]. For

a bridgeless PFC, the output voltage ground is always floating relative to the AC line input.

Thus, all parasitic capacitance including MOSFET drain to earth and the output terminals to the

earth ground contribute to common mode noise. This large dv/dt at each phase's switching node

leads to an increased common mode noise that is difficult to filter. At the same time, the

16

switching node Q2 and D2 are directly connected to input line terminal, which leads to high

dv/dt common mode noise.

The semi-bridgeless PFC boost converter is proposed [60] to address the EMI noise issues

for BL PFC Converters by adding two slow diodes, Da and Db to the input line as shown in

Figure 1-11. The added diodes have no effect on the efficiency of converter since they are in

parallel with another semiconductor when they conduct. As a matter of fact, they will reduce the

current stress on the min MOSFET in BL PFC configuration.

Figure 1-11: Bridgeless PFC boost topology

1.3.4. Interleaved PFC Boost Converter

The interleaved PFC boost converter (IL PFC) illustrated in Figure 1-12 is simply two

boost converters in parallel operating 180° out of phase. The input current is the sum of the two

inductor currents IL1 and IL2. Because the inductor’s ripple currents are out of phase, they tend to

cancel each other and reduce the input ripple current caused by the boost inductors [61-67].

The maximum input inductor ripple current cancellation occurs at 50% duty cycle. The output

capacitor current is the sum of the two boost diode currents less the dc output current.

Interleaving reduces the output capacitor ripple current as a function of the duty cycle [64]. As

the duty cycle approaches 0%, 50%, and 100% duty cycle, the sum of the two diode currents

approaches dc. At these points, the output capacitor only has to filter the inductor ripple current.

17

In order to design the IL PFC converter, it should be treated as two conventional boost PFC

converters with half of power rating. Therefore, all equations for the inductor, transistor, and

diode in conventional PFC are valid here, since the stresses are unchanged except the ripple

current through output capacitors.

The input bridge diode has the same power rating as the conventional PFC boost converter. But

the capacitor will get the most benefit of interleaving through reduced current ripple.

Figure 1-12: Interleaved PFC boost topology

As it can be noted, the input rectifier current is exactly the same as input rectifier current

in a conventional boost PFC converter, but the inductor current is exactly half, and boost diode

and boost MOSFET have less stress, due to the fact that they have to deliver half of the power as

in a conventional boost PFC converter. The capacitors will get the most benefits, due to

interleaving of each converter. In addition, the interleaved boost converter takes advantage of

paralleling semiconductors, and by having them switched out of phase, it doubles the effective

switching frequency and introduces smaller input current ripples, so the input EMI filters will be

smaller [68, 69]. But it still has the problem of heat management for the input rectifier diode

bridges, and suffers from lower efficiency at light load, and low line conditions.

18

1.3.4.1. PSIM Simulation of Interleaved Boost Converter

PSIM simulation software was used to verify steady state waveforms of each component.

Figure 1-13 shows the PSIM simulation circuits of the bridgeless boost PFC converter.

Figure 1-13: PSIM simulation circuit for interleaved PFC boost converter: Vin = 240V, Vo = 400V, Po =

3400W and fsw = 70 kHz

As it can be seen, the power stage section of the converter consists of two boost inductors,

Ld1 and Ld2, two fast boost diodes, Db1 and Db2, two MOSFETs, MOS1 and MOS2 and their

19

body diodes. Also it consists of two current loops and one voltage loop. The sensed input voltage

is multiplied by the compensated output voltage, and then generates a control signal to be

compared with the switching carrier waveforms to generate the gating signals for main

MOSFETs.

Figure 1-14 shows the PSIM simulation results of an interleaved PFC boost converter.

The input current is in phase with input voltage, and it has close to unity power factor. Also the

output voltage is regulated at around 400V, with a 120 Hz low frequency ripple. The converter is

operating at 70 kHz switching frequency, 240 V input voltage and 3.4 kW output power.

Figure 1-14: Simulation waveforms for interleaved PFC boost converter including output voltage, input

voltage and input current: Vin = 240V, Vo = 400V, Po = 3400W and fsw = 70 kHz

1.3.5. Ripple Steering Technique in PFC Applications

Coupled magnetics filter techniques, known also as ripple steering techniques [70], have

existed for many years, and have been applied to different topologies in industries [71, 72]. As

shown in Figure 1-15, this technique replaces a series smoothing inductor with a pair of coupled

20

inductors and a blocking capacitor. Recent attention is given to the application of ripple steering

to PFC boost converters [73, 74]. Figure 1-16 shows the modified PFC boost converter with a

coupled inductor. Ripple-steering technique has several advantages in a PFC boost converter.

Since it eliminates most of the differential-mode conducted noise, it enables the reduction in

EMI filter size and complexity, especially in its differential filtering section (Cx capacitors and

differential mode inductors). Reducing Cx capacitors to a minimum has an additional benefit for

applications with tight specifications on standby consumption. Cx capacitors cause a

considerable reactive current to flow through the filter, which is a source of additional and

unwanted loss. Furthermore, the discharge resistor that must be placed in parallel to Cx for safety

can be higher. As a result, both losses will be minimized. Although the control strategy of the

PFC stage is similar to that of a conventional boost converter, but their power stage transfer

functions are different. No modeling has been done to verify the effect of added coupled filter to

the power stage transfer functions, and thus the design of the feedback loop compensator.

Figure 1-15: Smoothing transformer in a coupled filter

Figure 1-16: Modified PFC boost converter with coupled inductors

21

Replacing the inductor in a conventional boost converter shown in Figure 1-3 with the

coupled inductor and a blocking capacitor shown in Figure 1-15 will result to the modified boost

converter with coupled magnetic shown in Figure 1-16.

1.3.6. Average Switch Model

Models for the PWM-switch were first introduced in [75, 76]. It was then adopted for

different converters in [77-81]. For the average switch model of a converter, the active and

passive switches shown in Figure 1-17 are replaced with the PWM-switch equivalent circuits

shown in Figure 1-18 and Figure 1-19.

Figure 1-17: One active switch and one passive switch

Figure 1-18: PWM-switch

Figure 1-19: Averaged model of PWM-switch

22

Using this model and applying that to modified boost converter with coupled inductors,

any of the transfer functions between the output variables (output voltage and inductor current)

and the input variables (input voltage and duty ration) can be derived.

1.4. Motivation and Objectives of the Thesis

The objective of this thesis is to investigate and conduct research on the high-

performance single-phase solutions for AC-DC power factor corrected converters for plug-in

hybrid electric vehicle battery chargers.

A variety of circuit topologies and control methods have been developed for PFC

applications. The two-stage approach is the proper candidate for PHEV battery chargers, where

the power rating is relatively high, and lithium-ion batteries are used as the main energy storage

system. The improvement of overall charger efficiency is critical for the emergence and

acceptance of these vehicular technologies, as the charger efficiency increases, the charge time

and utility cost decreases. Additionally, to meet the efficiency and power factor requirements and

regulatory standards for the AC supply mains, power factor correction is essential.

Due to limited space in vehicle and increasing power consumption, chargers are required

to deliver more power with smaller volume. As a key component of a charger system, the front-

end AC/DC converter must achieve high efficiency and power density. In this dissertation,

several conventional plug in hybrid electric vehicle charger front end AC-DC converter

topologies are investigated and a new bridgeless interleaved and a phase shifted semi-bridgeless

power factor corrected converter are proposed to improve the efficiency and performance, which

is critical to minimize the charger size, charging time and the amount and cost of electricity

drawn from the utility.

23

1.5. Outline of the Thesis

This thesis consists of five chapters. Chapter 1 introduces the subject of AC-DC power

factor corrected boost converters with particular focus on its application for plug-in hybrid

electric vehicles. This chapter establishes motivation and sets objectives for the research

contributions presented in Chapters 2-5. A literature review on existing topologies -

conventional boost converter, bridgeless boost converter and interleaved boost converter - modes

of operation, control techniques and PSIM simulation of existing topologies are presented in

Chapter 1. Chapter 2 proposes a new bridgeless interleaved PFC boost converter for PHEV

battery charger applications. The circuit description and steady state analysis, semiconductor

loss analysis, design procedure, logic implementation, PSIM simulation and experimental results

are presented. Chapter 3 introduces a new phase shifted semi-bridgeless PFC boost converter for

PHEV battery charger applications. The circuit description and steady state analysis,

semiconductor loss analysis, design procedure, logic implementation, PSIM simulation and

experimental results are presented. A new average switch modeling for ripple steering technique

applied to PFC boost converters and compensation network design based on derived voltage and

current power stage transfer functions are presented in Chapter 4. Simulation and experimental

results are presented to verify the model. Chapter 5 summarizes the contributions of the research

presented in this thesis and gives recommendations for future work.

24

CHAPTER 2. Bridgeless Interleaved PFC Boost Converter

2.1. Introduction

In this chapter, a new bridgeless interleaved power factor corrected converter for plug in

hybrid electric vehicle charger front end AC-DC converter topology is investigated and proposed

to improve the efficiency and performance, which is critical to minimize the charger size,

charging time, and the amount and the cost of electricity drawn from the utility. A detailed

analytical model for this topology is developed, enabling the calculation of power losses and

efficiency. Experimental and simulation results of a prototype boost converter converting

universal AC input voltage to 400 V DC at 3.4 kW are given to verify the proof of concept, and

analytical work reported in this chapter. The results show a power factor of greater than 0.99

from 750 W to 3.4 kW, THD less than 5% from half load to full load, and a peak efficiency of

98.94 % at 265 V input, and 1.2 kW output power.

2.2. Circuit Description and Steady State Analysis

The bridgeless interleaved (BLIL) PFC converter [82] shown in Figure 2-1 is proposed to

address the problems discussed in Chapter 1 for the conventional boost, bridgeless boost, and

interleaved boost topologies. This converter introduces two more MOSFETs and two more fast

diodes in place of four slow diodes used in the input bridge of the benchmark interleaved boost

PFC converter. To analyze the circuit operation, the input line cycle has been separated into the

positive and negative half cycles, as explained in sub-sections that follow. In addition, the

detailed circuit operation depends on the duty cycle, therefore positive half cycle operation

analysis is provided for D > 0.5 and D < 0.5.

25

Figure 2-1: Proposed bridgeless interleaved (BLIL) PFC boost converter

2.2.1. Positive Half Cycle Operation

Referring to Figure 2-1, during the positive half cycle, when the AC input voltage is positive,

Q1/Q2 turn on and current flows through L1 and Q1 and continues through Q2 and then L2,

returning to the line while storing energy in L1 and L2. When Q1/Q2 turn off, energy stored in

L1 and L2 is released as current flows through D1, through the load and returns through the body

diode of Q2 back to the input mains.

With interleaving, the same mode happens for Q3/Q4, but with a 180 degree phase delay.

The operation for this mode is Q3/Q4 on storing energy in L3/L4 through the path L3-Q3-Q4-L4

back to the input. When Q3/Q4 turn off, energy is released through D3 to the load and returning

through the body diode of Q4 back to the input mains.

2.2.2. Negative Half Cycle Operation

Referring to Figure 2-1, during the negative half cycle, when the AC input voltage is negative,

Q1/Q2 turn on and current flows through L2 and Q2 and continues through Q1 and then L1,

returning to the line while storing energy in L2 and L1. When Q1/Q2 turn off, energy stored in

L2 and L1 is released as current flows through D2, through the load and returns through the body

diode of Q1 back to the input mains.

26

With interleaving, the same mode happens for Q3/Q4, but with a 180 degree phase delay. The

operation for this mode is Q3/Q4 on storing energy in L3/L4 through the path L4-Q4-Q3-L3

back to the input.

2.2.3. Detailed Positive Half Cycle Operation and Analysis for D > 0.5

The detailed operation of the proposed BLIL PFC converter depends on the duty cycle. During

any half cycle, the converter duty cycle is either greater than 0.5 (when the input voltage is

smaller than half of output voltage) or smaller than 0.5 (when the input voltage is greater than

half of output voltage).

Figure 2-2: Interval 1: Q1 and Q2 are “ON”, and body diode of Q4 conducting

Figure 2-3: Interval 2 and 4: Q1, Q2, Q3 and Q4 are “ON”

27

Figure 2-4: Interval 3: Q3 and Q4 are “ON”, and body diode of Q2 conducting

Figures 2-2 to 2-4 show the three unique operating interval circuits of the proposed converter

for duty cycles greater than 0.5 during positive half cycle operation. Waveforms of the proposed

converter during these conditions are shown in Figure 2-5.

Since the switching frequency of proposed converter is much higher than the frequency of input

line voltage, the input voltage v2 is considered constant during one switching period T4. The

input voltage is given by:

56 √2 9:$;< % 2-1 In a positive half cycle of the input voltage, the duty ratio of the proposed converter determines

the following voltage relation:

=>? &&,0 2-2 The intervals of operation are explained as follows. In addition, the ripple current components

are derived, enabling calculation of the input ripple current, which provides design guidance to

meet the required input current ripple standard.

Interval 1 [t0-t1]: At t0, Q1/ Q2 are ON, and Q3/Q4 are off, as shown in Figure 2-2. During

this interval, the current in series inductances L1 and L2 increases linearly and stores the energy

in these inductors. The ripple currents in Q1 and Q2 are the same as the current in series

28

inductances L1 and L2, where the ripple current is given by:

Figure 2-5: BLIL PFC boost converter steady-state waveforms at D > 0.5

29

∆;A& &A -A+ 561 − )': 2-3 The current in series inductances L3 and L4 decreases linearly and transfers the energy to the

load through D3, Co and body diode of Q4. The ripple current in series inductances L3 and L4

is given by:

∆;AD &AE-AF 9 − 561 − )': 2-4 The input ripple current is the sum of currents in L1/L2 and L3/L4:

∆G6 &A -A+ 9 1 − )': 2-5 Interval 2 [t1-t2]: At t1, Q3/Q4 are turned on, while Q1/Q2 remain on, as shown in Figure

2-3. During this interval, the current in the four inductors each increase linearly, storing energy

in these inductors. The ripple currents in Q1 and Q2 are the same as the ripple current in series

inductances L1 and L2 as given by:

∆;A& &A -A+ 56) − &H': 2-6 Similarly, the ripple currents in Q3 and Q4 are the same as the ripple current in series

inductances L3 and L4:

∆;AD &AE-AF 56) − &H': 2-7 The input ripple current is the sum of currents in L1/L2 and L3/L4:

∆G6 HA -A+ 56) − &H': 2-8 Interval 3 [t2-t3]: At t2, Q1/Q2 are turned off, while Q3/ Q4 remain on, as shown in

Figure 2-4. During this interval, the current in series inductances L3 and L4 increases linearly

and stores the energy in these inductors. The ripple currents in Q3 and Q4 are the same as the

ripple current in series inductances L3 and L4:

30

∆;AD &AE-AF 561 − )': 2-9 The current in L1 and L2 decreases linearly and transfers the energy to the load through D1, Co

and body diode of Q2. The ripple current in series inductances L1 and L2 is given by:

∆;A& &A -A+ 9 − 561 − )': 2-10 The input ripple current is the sum of currents in L1/L2 and L3/L4:

∆G6 &A -A+ 9 1 − )': 2-11 Interval 4 [t3-t4]: At t3, Q3/Q4 remain on, while Q1/Q2 are turned on, as shown in Figure

2-3. During this interval, the currents in the four inductors each increase linearly, storing energy

in these inductors. The ripple currents in Q1 and Q2 are the same as the ripple currents in L1/L2:

∆;A& &A -A+ 56) − &H': 2-12 Similarly, the ripple currents Q3 and Q4 are the same as the ripple current in series inductances

L3 and L4:

∆;AD &AE-AF 56) − &H': 2-13 The input ripple current is the sum of currents in L1/L2 and L3/L4:

∆G6 HA -A+ 56) − &H': 2-14 2.2.4. Detailed Positive Half Cycle Operation and Analysis for D < 0.5

Figure 2-6 to Figure 2-8 show the operating interval circuits of the proposed converter for duty

cycles smaller than 0.5 during the positive half cycle. The waveforms of the proposed converter

during these conditions are shown in Figure 2-9. The intervals of operation are explained as

follows:

31

D1L1 D2 D3 D4

Q1 Q2 Q3 Q4

L3

L2

L4

Co

L

O

A

D

Vin

Vg1 Vg2

Figure 2-6: Interval 1 and 3: Body diodes of Q2 and Q4 conducting

Figure 2-7: Interval 2: Q1 and Q2 are “ON”, and body diode of Q4 conducting

Figure 2-8: Interval 4: Q3 and Q4 are “ON”, and body diode of Q2 conducting

32

Interval 1 [t0-t1]: At t0, Q1 and Q2 turn off, while Q3 and Q4 remain off, as shown in

Figure 2-6. During this interval, the current in series inductances L1 and L2 decreases linearly

and transfers the energy to the load through D1, Co and body diode of Q2. The ripple current in

series inductances L1 and L2 is:

∆;A& &A -A+ 9 − 56&H − )': 2-15 In addition, the current in the series inductances L3 and L4 also decreases linearly, transferring

the energy to the load through D3, Co and body diode of Q4. The ripple current in series

inductances L3 and L4 is:

∆;AD &AE-AF 9 − 56&H − )': 2-16 The input current is the sum of currents in L1/L2 and L3/L4:

∆G6 HA -A+ 9 − 56&H − )': 2-17 Interval 2 [t1-t2]: At t1, Q1/Q2 turn on, while Q3/Q4 remain off, as shown in Figure 2-7.

During this interval, the current in series inductances L1 and L2 increases linearly, storing

energy in these inductors. The ripple currents in Q1 and Q2 are the same as the current in series

inductances L1 and L2, where the ripple current is given by:

∆;A& &A -A+ 56)': 2-18 The current in series inductances L3 and L4 decreases linearly and transfers the energy to the

load through D3, Co and body diode of Q4. The ripple current in L3 and L4 is:

∆;AD &AE-AF 9 − 56)': 2-19 The input ripple current is the sum of the currents in L1/L2 and L3/L4:

∆G6 &A -A+ 9 )': 2-20

33

t

t

t

t

t

t

Vg2

t

Vg1

Input Current

DTS(1-D)TS

t0 t1 t2 t3 t4 t5

∆Iin

Gating Signals

Inductor Current

FET Current

Boost Diode

Current

Body Diode

Currentt

IL1

IQ1

ID1

IQd3

IL3

IQ3

ID3

IQd1

Figure 2-9: BLIL PFC boost converter steady-state waveforms at D < 0.5

34

Interval 3 [t2-t3]: At t2, Q1/Q2 are turned off, while Q3/Q4 remain off, as shown in

Figure 2-6. During this interval, the current in series inductances L1 and L2 decreases linearly

and transfers the energy to the load through D1, Co and body diode of Q2. The ripple current in

series inductances L1 and L2 is given by:

∆;A& &A -A+ 9 − 56&H − )': 2-21

Similarly, the current in the series inductances L3 and L4 also decreases linearly, transferring the

energy to the load through D3, Co and body diode of Q4. The ripple current in series inductances

L3 and L4 is:

∆;AD &AE-AF 9 − 56&H − )': 2-22

The input current is the sum of currents in L1/L2 and L3/L4:

∆G6 HA -A+ 9 − 56&H − )': 2-23

Interval 4 [t3-t4]: At t3, Q3/Q4 are turned on, while Q1/Q2 remain off, as shown in Figure

2-8. During this interval, the current in series inductances L3 and L4 increases linearly and stores

the energy in these inductors. The ripple currents in Q3 and Q4 are the same as the current in

series inductances L3 and L4, where the ripple current is given by:

∆;AD &AE-AF 56)': 2-24

The current in series inductances L1 and L2 decreases linearly and transfers the energy to the

load through D2, Co and body diode of Q4. The ripple current in L1 and L2 is:

∆;A& &A -A+ 9 − 56)': 2-25

The input ripple current is the sum of currents in L1/L2 and L3/L4:

∆G6 &A -A+ 9 )': 2-26

35

The operation of converter during the negative input voltage half cycle is similar to the operation

of converter during the positive input voltage half cycle.

2.3. Loss Modeling

In order to properly select the power stage components of a converter and calculate the

associated power losses, it is necessary to determine the RMS and the average values of their

currents [83]. In a typical boost converter, the MOSFET and diode current waveforms are

pulsed-width modulated, with both the duty cycle and peak amplitude varying with the AC input.

Without an effective mathematical method for computing these RMS and average values, the

proper design and selection of power stage components can be flawed. The following

assumptions were made in order to analyze the converters and to derive the stress equations:

a) These calculations are based on CCM operation of the PFC boost converter.

b) Assuming unity power factor, the line current is in phase and shape with the input line

voltage – a sinusoidal waveform.

c) The PFC output voltage is DC with no voltage ripple.

In a typical boost converter, the converter MOSFET duty cycle is given by:

IJ% 1 − |?L|= 1 − MN|:6 | = 2-27 Therefore

IJ,O: *&P Q R1 − MN|:6 | = SHPT U% 2-28 Assuming the inductor current is a sinusoidal waveform:

;A % GV|$;< %| 2-29 The instantaneous MOSFET current and its RMS current can be derived respectively:

;J% GV|$;<%|. IJ,O: 2-30

36

GJ,O: *&P Q RGV|$;<%|1 − MN|:6 |= SHPT U% 2-31 The inductor current ripple is assumed to be half of peak inductor current:

∆G &H MNH 2-32 The high frequency ripple components of inductor current is assumed to be a triangular

waveform with a fixed duty cycle, so the RMS current in each inductor is defined by:

GA,O: * &√H MNH H + &H√D ∆G H YZ√D ?LMN 2-33 The boost diode duty cycle is given by:

I0% 1 − IJ% MN|:6 |= 2-34 Therefore the instantaneous boost diode current and its RMS current can be derived respectively:

;0% GV|$;<%| MN|:6 |= 2-35 G0,O: *&P Q RGV|$;<%|MN|:6 |= SHPT U% 2-36

The output capacitor current has high frequency and low frequency components. The low

frequency component is simply calculated by:

G[,O:A\ =√H √HH == 2-37 The high frequency RMS ripple current component is:

G[,O:/\ ?L= * &] =]P MN − =+?L+ 2-38 The same method was used to derive RMS current in different topologies. Table 2-1

shows a summary of component RMS current stress for conventional boost converter and

bridgeless boost converter. Table 2-2 provides the same summary for interleaved boost converter

and bridgeless interleaved boost converter.

37

Table 2-1: Summary of component RMS current for conventional boost and bridgeless boost topologies

Topology Conventional PFC Bridgeless PFC

Boost Inductor ^9748 69V ^9748 69V

Input Bridge Diode 69V Not Applicable

Boost Fast Diode ^32 69 √32 69

Boost Transistor 6√69V9^3e39VH + 49H − 649V9e

6√69V9^3e39VH + 49H − 649V9e

Boost Transistor Intrinsic

Diode Not Applicable √32 69

Output Capacitor Ripple (LF) √22 9 √22 9

Output Capacitor Ripple (HF) √229 *36H − 2H √229 *36H − 2H

Table 2-2: Summary of component RMS current for interleaved and bridgeless interleaved boost topologies

Topology Interleaved PFC Bridgeless Interleaved PFC

Boost Inductor 54√3 69V 54√3 69V Input Bridge Diode 69V

Not Applicable

Boost Fast Diode ^32 62. 9 √34 69

Boost Transistor 62√69V9^3e39VH + 49H − 649V9e

62√69V9^3e39VH + 49H − 649V9e

Boost Transistor Intrinsic

Diode Not Applicable √34 69

Output Capacitor Ripple (LF) √22 9 √22 9

Output Capacitor Ripple (HF) 69 ^ 16 96e 9V − H6H 69 ^ 16 96e 9V − H6H

38

Table 2-3 shows a summary of component Average current stress for conventional boost

converter and bridgeless boost converter. Table 2-4 provides the same summary for interleaved

boost converter and bridgeless interleaved boost converter.

Table 2-3: Summary of component average current for conventional boost and bridgeless boost topologies

Topology Conventional PFC Bridgeless PFC

Boost Inductor 4e 69V 0

Input Bridge Diode 2e 69V Not Applicable

Boost Fast Diode 69 12 69

Boost Transistor 64e 19V − 19 12 69

Boost Transistor Intrinsic

Diode Not Applicable 12 69

Table 2-4: Summary of component average current for interleaved and bridgeless interleaved boost

topologies

Topology Interleaved PFC Bridgeless Interleaved PFC

Boost Inductor 2e 69V 0 Input Bridge Diode 2e 69V

Not Applicable

Boost Fast Diode 62. 9 14 69

Boost Transistor 64e 19V − 12 19 14 69

Boost Transistor Intrinsic

Diode Not Applicable 14 69

39

As can be noted, in both the bridgeless topology and bridgeless interleaved topology, a

new loss has been introduced in the intrinsic body diodes of the MOSFETs, but since input

bridge rectifiers were eliminated, there is some efficiency gain in overall performance of these

topologies. The intrinsic body diode of MOSFETs conduct when the boost transistors are off and

its value is the same as the current in the boost diodes, when they conduct and transfer energy to

the output capacitors.

Also the low frequency RMS ripple current through output capacitors is constant and

interleaving has no effect on it. But the high frequency ripple current will be reduced

significantly, as it is shown in Figure 2-10. Also it is noted that as the input voltage increases, the

high frequency ripple reduces.

Figure 2-10: RMS ripple current through output capacitors vs input voltage at Po = 3400 W, fsw = 70 kHz

Figure 2-11 shows the loss distribution of the semiconductors in the four topologies

investigated in this thesis at Vin = 240 V, Po = 3400 W, Vo = 400 V and fsw = 70 kHz. The regular

diode losses are only conduction losses in bridge rectifier diodes. Because of low reverse

40

recovery characteristics of SiC diodes, it is the selected component for boost diodes; therefore it

is assumed that there is only conduction loss in fast diodes. Switching losses, conduction losses,

gate charge losses and ½ CV2 losses are all included in MOSFET losses. It should be noted that

the inductor loss analysis is not included in this study.

Figure 2-11: Comparison of the estimated loss distribution in the semiconductors at 70kHz switching

frequency, 240V input, 3400W load at 400V

The regular diodes in the input bridge rectifiers have the largest share of losses among the

topologies with the input bridge rectifier. The bridgeless topologies eliminate this large loss

component (~27.6 W). However, the tradeoff is that the MOSFET losses are higher and the

intrinsic body diodes of the MOSFETs conduct, producing new losses (~7.85 W). The fast

diodes in the bridgeless interleaved PFC have slightly lower power losses, since the boost diode

average current is lower in these topologies. Overall, the MOSFETs are under more stress in

bridgeless topologies, but the total semiconductor losses for the proposed bridgeless interleaved

27.6

W

12.9

W

8.3

W

0 W

48.7

W

0 W

12.7

W

19.1

W

7.8

W

39.6

W

27.6

W

12.9

W

8.3

W

0 W

48.7

W

0 W

11.3

W 16.6

W

7.8

W

35.7

W

0

10

20

30

40

50

60

Reg

ula

r

Dio

des

Fast

Dio

des

FE

Ts

Intr

insi

c

Bod

y D

iod

es

Tota

l L

oss

es

Pow

er

Loss

es (

W)

Devices / Total Losses

Conventional Boost

Bridgeless Boost

Interleaved Boost

Bridgeless Interleaved Boost

41

boost are 37% lower than the benchmark conventional boost, 10% lower than the bridgeless

boost and 37% lower than the benchmark interleaved boost.

Since the bridge rectifier losses are so large, it was expected that the bridgeless interleaved

boost converter would have the least power losses among all four introduced topologies. Also it

was noted that the losses in the input bridge rectifiers were 56% of total losses in the

conventional PFC converter and in the benchmark interleaved PFC converter. Therefore

eliminating the input bridges in PFC converters is justified despite the fact that new losses are

introduced.

2.4. Math Modeling Verification

In order to verify the accuracy of converters modeled in section 2.3, the math modeling results

were compared with PSIM simulations for two different topologies. The RMS inductor current,

the RMS MOSFET current, the average boost diode current and the average MOSFET body

diode current for an interleaved boost PFC topology and a bridgeless boost PFC topology were

examined.

Figure 2-12 to Figure 2-14 show the PSIM simulation and Math modeled values of boost

inductor RMS current, MOSFET RMS current and boost diode average current for an interleaved

boost PFC converter at 240 V input and 120 V input voltages.

Figure 2-15 to Figure 2-18 show the PSIM simulation and Math modeled values of boost

inductor RMS current, MOSFET RMS current, boost diode average current and MOSFET body

diode average current for a bridgeless boost PFC converter at 240 V input and 120 V input

voltages.

42

Figure 2-12: Simulation and math modeling of inductor RMS current for an interleaved PFC converter

Figure 2-13: Simulation and math modeling of MOSFET RMS current for an interleaved PFC converter

0

1

2

3

4

5

6

7

8

9

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

40

00

Ind

uct

or

RM

S C

urr

en

t (A

)

Output Power (W)

Math Modelling - Vin = 240 V

PSIM Simulation - Vin = 240 V

Math Modelling - Vin = 120 V

PSIM Simulation - Vin = 120 V

0

1

2

3

4

5

6

7

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

40

00

MO

SF

ET

RM

S C

urr

en

t (A

)

Output Power (W)

Math Modelling - Vin = 240 V

PSIM Simulation - Vin = 240 V

Math Modelling - Vin = 120 V

PSIM Simulation - Vin = 120 V

43

Figure 2-14: Simulation and math modeling of boost diode average current for an interleaved PFC converter

Figure 2-15: Simulation and math modeling of inductor RMS current for a bridgeless PFC converter

0

0.5

1

1.5

2

2.5

3

3.5

4

4.5

5

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

40

00

Bo

ost

Dio

de

Av

era

ge

Cu

rre

nt

(A)

Output Power (W)

Math Modelling - Vin = 240 V

PSIM Simulation - Vin = 240 V

Math Modelling - Vin = 120 V

PSIM Simulation - Vin = 120 V

0

2

4

6

8

10

12

14

16

18

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

40

00

Ind

uct

or

RM

S C

urr

en

t (A

)

Output Power (W)

Math Modelling - Vin = 240 V

PSIM Simulation - Vin = 240 V

Math Modelling - Vin = 120 V

PSIM Simulation - Vin = 120 V

44

Figure 2-16: Simulation and math modeling of MOSFET RMS current for a bridgeless PFC converter

Figure 2-17: Simulation and math modeling of boost diode average current for a bridgeless PFC converter

0

1

2

3

4

5

6

7

8

9

10

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

40

00

MO

SF

ET

RM

S C

urr

en

t (A

)

Output Power (W)

Math Modelling - Vin = 240 V

PSIM Simulation - Vin = 240 V

Math Modelling - Vin = 120 V

PSIM Simulation - Vin = 120 V

0

0.5

1

1.5

2

2.5

3

3.5

4

4.5

5

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

40

00

Bo

ost

Dio

de

Av

era

ge

Cu

rre

nt

(A)

Output Power (W)

Math Modelling - Vin = 240 V

PSIM Simulation - Vin = 240 V

Math Modelling - Vin = 120 V

PSIM Simulation - Vin = 120 V

45

Figure 2-18: Simulation and math modeling of MOSFET body diode average current for a bridgeless PFC

2.5. Simulation Results

PSIM simulation software was used to verify steady state waveforms of each component.

Figure 2-19 shows the PSIM simulation circuits of the proposed BLIL PFC converter. As it can

be seen, the power stage section of converter consists of four boost inductors, Ld1 to Ld4, four

fast boost diodes, Db1 to Db4, four switches, Q1 to Q4 and their body diodes Dq1 to Dq4. Also

it consists of two current loops and one voltage loop. The sensed input voltage is multiplied by

the compensated output voltage, and then generates a control signal to be compared with the

switching carrier waveforms to generate the gating signals for the main MOSFETs.

0

1

2

3

4

5

6

7

8

9

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

40

00

MO

SF

ET

Bo

dy

Dio

de

Av

era

ge

Cu

rre

nt

(A)

Output Power (W)

Math Modelling - Vin = 240 V

PSIM Simulation - Vin = 240 V

Math Modelling - Vin = 120 V

PSIM Simulation - Vin = 120 V

46

Figure 2-19: PSIM simulation circuit for the proposed BLIL PFC boost converter: Vin = 240V, Vo = 400V,

Po = 3400W and fsw = 70 kHz

Figure 2-20 shows the PSIM simulation results of a BLIL PFC boost converter. The input

current is in phase with input voltage, and it has close to unity power factor. Also the output

voltage is regulated at around 400V, with a 120 Hz low frequency ripple. The converter is

operating at 70 kHz switching frequency, 240 V input voltage and 3.4 kW output power.

47

Figure 2-20: Simulation waveforms for the proposed BLIL PFC boost converter including output voltage,

input voltage and input current: Vin = 240V, Vo = 400V, Po = 3400W and fsw = 70 kHz

2.6. Logic Implementation

A standard two-phase interleaved CCM PFC controller from Texas Instrument, UCC28070

was used to implement the logic circuitry of the prototype unit. The key features of this

controller are as following:

• Interleaved average current mode PWM control

• Advanced current synthesizer for superior efficiency, accurate current sensing, and high

power factor

• Highly linear multiplier output with internal voltage feed-forward correction for near unity

power factor

• Programmable switching frequency (30 kHz to 300 kHz)

• Selectable frequency dithering for reduced EMI

48

• Phase management for high-efficiency light-load operation

A simplified logic implementation of this controller applied to bridgeless interleaved PFC

boost converter is given in Figure 2-21.

Figure 2-21: Logic implementation of BLIL PFC boost converter using UCC28070 controller

49

2.7. Experimental Results

An experimental prototype, illustrated in Figure 2-22, was built to verify the operation of the

proposed converter. Figure 2-23 shows the input voltage, input current and PFC bus voltage of

the converter under the following test conditions: Vin = 240 V, Iin = 15 A, Po = 3400 W, Vo =

400 V, fsw = 70 kHz. The input current is in line and phase with the input voltage, and its shape

is close to a sinusoidal waveform. Table 2-3 shows the semiconductors and power components

used in the 3.4 kW CCM experimental prototype, a bridgeless interleaved PFC converter.

Figure 2-22: Breadboard prototype of BLIL PFC boost converter

50

Figure 2-23: Breadboard proposed BLIL PFC experimental waveforms;

Test condition: Po = 3400W, Vin = 240V, Iin = 15A.

Figure 2-24: Gating signal, boost MOSFET and diode current for D > 0.5

Input Voltage

Ch2 = Vin 100V/div.

Output Voltage Ch1= Vo 100V/div.

Input Current

Ch3 = Iin 10A/div.

Diode Current

Ch3 = ID1 2A/div.

Gating Signal

Ch1 = Vg 10V/div.

MOSFET Current

Ch4 = IQ1 2A/div.

51

Figure 2-25: Gating signal, boost MOSFET and diode current for D < 0.5

Figure 2-26: Load transient response from FL to NL (3400W to 0W)

Diode Current

Ch3 = ID1 5A/div.

Gating Signal

Ch1= Vg 10V/div.

MOSFET Current

Ch4 = IQ1 5A/div.

Input Voltage

Ch2 = Vin 100V/div.

Output Voltage Ch1= Vo 100V/div.

Input Current

Ch4 = Iin 10A/div.

Output Current

Ch3 = Iin 5A/div.

52

Figure 2-27: Load transient response from NL to FL (0W to 3400 W)

Table 2-5: Component / devices used in prototype unit

Topology Device Part # / Value # of devices

Ben

chm

ark I

nte

rlea

ved

PF

C Regular Diode 25ETS08S 4

Fast Diode IDB06S60C 2

MOSFET IPB60R099CP 2

Inductor Kool mu 77071 core

60 turns / 400 µH 2

Capacitor EKXJ451ELL820 / 82 µF 10

Bri

dgel

ess

Inte

rlea

ved

PF

C

Fast Diode CSD10060 4

MOSFET IPP60R099CP 4

Inductor Kool mu 77071 core

60 turns / 400 µH 4

Capacitor EKXJ451ELL820 / 82 µF 10

Input Voltage

Ch2 = Vin 100V/div. Output Voltage

Ch1= Vo 100V/div.

Input Current

Ch4 = Iin 10A/div.

Output Current

Ch3 = Iin 5A/div.

53

In order to verify the quality of the input current, its harmonics, up to 39th harmonic, are given

and compared with the IEC 61000-3-2 standard.

Figure 2-28: Input current harmonics at full load for Vin = 120 V at Po = 1700 W and 240 V at Po = 3400 W

Figure 2-29: RMS Total harmonics distortion vs. output power at Vin = 120V and Vin = 240V

0

0.5

1

1.5

2

2.5

3 5 7 9

11

13

15

17

19

21

23

25

27

29

31

33

35

37

39

Ha

rmo

nic

Cu

rre

nt

(A)

Harmonics Order

EN 61000-3-2 Class D Limits (A)

Amplitude (A) Vin = 120 V

Amplitude (A) Vin = 240 V

0

5

10

15

20

25

30

35

40

45

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

TH

D (

%)

Output Power (W)

Vin=240

Vin=120

54

Figure 2-30: Power factor vs. output power at Vin = 120V and Vin = 240V

Figure 2-31: Efficiency vs. output power for the proposed bridgeless interleaved PFC boost converter

0.84

0.86

0.88

0.9

0.92

0.94

0.96

0.98

1

1.02

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

Po

we

r fa

cto

r

Output Power (W)

Vin=240

Vin=120

93

94

95

96

97

98

99

100

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

40

00

45

00

50

00

Eff

icie

ncy

(%

)

Output Power (W)

Vin = 90 V

Vin = 120 V

Vin = 220 V

Vin = 240 V

Vin = 265 V

55

Figure 2-32: Efficiency vs. output power for different topologies at Vin = 240V

Figure 2-24 shows the gating signal, boost MOSFET current, and boost diode current for duty

cycles greater than 50%. Figure 2-25 shows the same voltage and current for duty cycles smaller

than 50%. Figure 2-26 and Figure 2-27 show the output voltage transient response to a change to

the load form FL to NL and vice versa. As can be noted, the output voltage regulates right away.

Figure 2-28 shows the input current harmonics versus harmonic numbers at full load for 120 V

and 240 V input voltages. It is clearly shown that the generated harmonics are well below IEC

61000-3-2 standard for the input line harmonics which is required for PHEV chargers.

In Figure 2-29, the input current total harmonics distortions are given at full load and for 120

V and 240 V input voltages. It can be noted that mains current THD are smaller than 5% from

50% load to full load and it is compliant to IEC 61000-3-2. Another parameter to show the

quality of the input current is the power factor. In Figure 2-30, the converter power factor is

shown over the entire load range for different input voltages. As it can be seen, the power factor

is greater than 0.99 from 50% load to full load.

94

95

96

97

98

99

100

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

40

00

45

00

Eff

icie

ncy

(%

)

Output Power (W)

Bridgeless Interleaved PFC Converter Vin = 240 V

Benchmark Interleaved PFC Converter Vin = 240 V

56

The efficiency of converter versus output power for different input voltages is provided in

Figure 2-31. With the proposed bridgeless interleaved PFC converter a peak efficiency of

98.94% was reached at 265 V input and 1.2 kW output power. High efficiency over the entire

load range is achieved in this topology, enabling fewer problems with heat dissipation and

cooling systems. Furthermore, a higher efficiency means more power is available from the mains

feed to charge the batteries, reducing charging time and electricity costs.

Figure 2-32 illustrates the measured experimental efficiency of benchmark interleaved

boost PFC and BLIL boost PFC topology discussed in this chapter for 240 V input voltage. The

proposed BLIL PFC converter has the highest efficiency over the entire load range.

Figure 2-33: Loss reduction as a function of output power at Vin = 240V, Vo = 400 V and 70 kHz for the

proposed BLIL converter compared to the benchmark interleaved boost converter

0

10

20

30

40

50

60

70

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

Loss

Re

du

ctio

n (

%)

Output Power (W)

57

Loss reduction curves as a function of output power for BLIL PFC against benchmark IL

PFC are provided in Figure 2-33 and Figure 2-34 for 240 V and 120 V.

Figure 2-34: Loss reduction as a function of output power at Vin = 120V, Vo = 400 V and 70 kHz for the

proposed BLIL converter compared to the benchmark interleaved boost converter

2.8. Conclusion

A high performance AC-DC boost converter topology has been presented in this chapter for

the front-end AC-DC converter in PHEV battery chargers. The proposed converter topology has

been analyzed and performance characteristics presented. Also an analytical model for the four

different topologies was developed, enabling the calculation of power losses and efficiencies. A

prototype converter was built to verify this proof-of-concept. The theoretical waveforms were

compared with the results taken from prototype unit. Also some key experimental waveforms

0

10

20

30

40

50

60

0

20

0

40

0

60

0

80

0

10

00

12

00

14

00

16

00

18

00

Loss

Re

du

ctio

n (

%)

Output Power (W)

58

were given. Finally, input current harmonics at each harmonic order were compared more

explicitly with the IEC 61000-3-2 standard limits. It can be noted that mains current THD are

smaller than 5% from 50% load to full load and the converter is compliant with the IEC 6100-3-

2 standard. The converter power factor was also provided for full power range and different input

voltages. It can be seen that the power factor is greater than 0.99 from 50% load to full load. The

proposed converter achieves a peak efficiency of 98.94 % at 265 V input and 1.2 kW output

power.

The converter topology shows a high input power factor, high efficiency over the entire

load range, and excellent input current harmonics. It is a potential option for single phase PFC in

higher power battery charging applications.

59

CHAPTER 3. Phase-Shifted Semi Bridgeless PFC Converter

3.1. Introduction

In this chapter, a phase shifted semi-bridgeless boost power factor corrected converter is

proposed for plug in hybrid electric vehicle battery chargers. The converter features high

efficiency at light loads and low lines, which is critical to minimize the charger size, charging

time, and the amount and cost of electricity drawn from the utility; the component count, which

reduces the charger cost; and reduced EMI. The converter is ideally suited for automotive level 1

residential charging applications.

A detailed converter description and steady state operation analysis of this converter is

presented. Experimental results of a prototype boost converter, converting universal AC input

voltage to 400 V DC at 3.4 kW are given and the results are compared to a benchmark

interleaved boost converter to verify the proof of concept, and analytical work reported. The

results show a power factor greater than 0.99 from 750 W to 3.4 kW, THD less than 5% from

half load to full load, a peak efficiency of 98.6 % at 240 V input, and 1000 W load.

3.2. Circuit Description and Steady State Analysis

The phase shifted semi-bridgeless topology [84], shown in Figure 3-1, is proposed as another

solution to address the problems outlined in chapter 1 for the conventional boost, bridgeless

boost and interleaved boost topologies. The proposed topology has high efficiency at light loads

and low lines, which is important to minimize the charger size, charging time, and the amount

and cost of electricity drawn from the utility; the component count, which reduces the charger

cost; and reduced EMI. The converter is ideally suited for automotive level 1 residential charging

applications in North America where the typical supply is limited to 120V and 1.44kVA.

60

Figure 3-1: Phase shifted semi-bridgeless PFC boost topology

The proposed topology introduces two more slow diodes (Da and Db) to the bridgeless

configuration to link the ground of the PFC to the input line. However, the current does not

always return through these diodes, so their associated conduction losses are low. This occurs

since the inductors exhibit low impedance at the line frequency, a large portion of the current

flows through the MOSFET intrinsic body diodes. Also the gating signals for MOSFETs are

180° out of phase. To analyze the circuit operation, the input line cycle has been separated into

the positive and negative half-cycles, as explained in the sub-sections that follow. In addition, the

detailed circuit operation depends on the duty cycle. Positive half-cycle operation analysis is

provided for D > 0.5 and D < 0.5.

3.2.1. Positive Half Cycle Operation

Referring to Figure 3-1, during the positive half-cycle, when the AC input voltage is

positive, Q1 turns on and current flows through L1 and Q1 and continues through Q2 and then

L2, returning to the line while storing energy in L1 and L2. When Q1 turns off, energy stored in

L1 and L2 is released as current flows through D1, through the load and returns through the body

diode of Q2/partially through Db back to the input.

61

3.2.2. Negative Half Cycle Operation

Referring to Figure 3-1, during the negative half-cycle, when the AC input voltage is

negative, Q2 turns on and current flows through L2 and Q2 and continues through Q1 and then

L1, returning to the line while storing energy in L2 and L1. When Q2 turns off, energy stored in

L2 and L1 is released as current flows through D2, through the load and returns split between the

body diode of Q1 and Da back to the input.

3.2.1. Detailed Positive Half Cycle Operation and Analysis for D > 0.5

The detailed operation of the proposed converter depends on the duty cycle. During any

half-cycle, the converter duty cycle is either greater than 0.5 (when the input voltage is smaller

than half of output voltage) or smaller than 0.5 (when the input voltage is greater than half of

output voltage). The three unique operating interval circuits of the proposed converter are

provided in Figure 3-2 to Figure 3-4 for duty cycles larger than 0.5 during the positive half-cycle.

Waveforms of the proposed converter during positive half-cycle operation with D>0.5 are

shown in Figure 3-5. The intervals of operation are explained as follows. In addition, the ripple

current components are derived, enabling calculation of the input ripple current, which provides

design guidance to meet the required input current ripple standard.

D1 D2

Q1 Q2

CO

L1

L2

L

O

A

D

Vin

Vg1 Vg2

DaDb

Figure 3-2: Interval 1and 3: Q1 and Q2 are ON

62

Figure 3-3: Interval 2: Q1 ON, body diode of Q2 conducting

Figure 3-4: Interval 4: Q1 OFF and Q2 ON

Interval 1 [t0-t1]: At t0, Q1/ Q2 are on, as shown in Figure 3-2. During this interval, the

current in series inductances L1 and L2 increases linearly and stores the energy in these

inductors. The energy stored in Co provides energy to the load. The ripple currents in Q1 and Q2

are the same as the current in series inductances L1 and L2, where the ripple current is given by:

∆G6 &A -A+ 56 ) − &H': 3-1

Interval 2 [t1-t2]: At t1, Q1 is on and Q2 is off, as shown in Figure 3-3. During this

interval, the current in series inductances L1 and L2 continues to increase linearly and store the

energy in these inductors. The energy stored in Co provides the load energy. The ripple currents

in Q1 and body diode of Q2 are the same as the current in series inductances L1 and L2, where

the ripple current is given by:

63

∆G6 &A -A+ 56 1 − )': 3-2

Interval 3 [t2-t3]: At t2, Q1/Q2 are on again, and interval 1 is repeated, as shown in Figure

3-2. During this interval, the current in series inductances L1 and L2 increases linearly and stores

the energy in these inductors. The ripple currents in Q1 and Q2 are the same as the ripple current

in series inductances L1 and L2, as shown in equation (3-1).

Interval 4 [t3-t4]: At t3, Q1 is off and Q2 is on, as shown in Figure 3-4. During this

interval, the energy stored in L1 and L2 is released to the output through L1, D1, Q2 and L2. The

ripple currents in D1 and Q2 are the same as the ripple currents in L1 and L2:

∆G6 &A -A+ 56 − 91 − )': 3-3

Figure 3-5: Phase shifted semi-bridgeless boost converter steady-state waveforms at D > 0.5

64

3.2.1. Detailed Positive Half Cycle Operation and Analysis for D < 0.5

The three unique operating interval circuits of the proposed converter are given in Figure

3-6 to Figure 3-8 for duty cycles smaller than 0.5 during the positive half-cycle. The waveforms

of the proposed converter during these conditions are shown in Figure 3-9. The intervals of

operation are explained as follows.

Figure 3-6: Interval 1and 3: Q1 and Q2 are OFF, body diode of Q2 conducting

Figure 3-7: Interval 2: Q1 ON, body diode of Q2 conducting

Figure 3-8: Interval 4: Q1 OFF and Q2 ON

65

Figure 3-9: Phase shifted semi-bridgeless boost converter steady-state waveforms at D < 0.5

Interval 1 [t0-t1]: At t0, Q1/ Q2 are off, as shown in Figure 3-6. During this interval, the

energy stored in L1 and L2 are released to the output through L1, D1, body diode of Q2 and L2.

The ripple currents in D1 and body diode of Q2 are the same as the ripple currents in L1 and L2:

∆G6 &A -A+ 56 − 9&H − )': 3-4 Interval 2 [t1-t2]: At t1, Q1 is on and Q2 is off, as shown in Figure 3-7. During this

interval, the current in series inductances L1 and L2 continues to increase linearly and store the

energy in these inductors. The energy stored in Co provides energy to the load. The ripple

66

currents in Q1 and the body diode of Q2 are the same as the current in series inductances L1 and

L2, where the ripple current is given by:

∆G6 &A -A+ 56 )': 3-5 Interval 3 [t2-t3]: At t2, Q1/Q2 are off again, and interval 1 is repeated, as shown in

Figure 3-6. During this interval, the current in series inductances L1 and L2 increases linearly

and stores the energy in these inductors. The ripple currents in D1 and body diode of Q2 are the

same as the ripple current in series inductances L1 and L2, as shown in equation (3-4).

Interval 4 [t3-t4]: At t3, Q1 is off and Q2 is on, as shown in Figure 3-8. During this

interval, the energy stored in L1 and L2 is released to the output through L1, D1, Q2 and L2. The

ripple currents in D1 and Q2 are the same as the ripple currents in L1 and L2:

∆G6 &A -A+ 56 − 9)': 3-6 The operation of converter during the negative input voltage half-cycle is similar to the operation

of converter during the positive input voltage half-cycle.

3.3. Loss Modeling

The estimated loss distribution of the semiconductors is provided in Figure 3-10 at 70 kHz

switching frequency, 240V input and 3300W load for benchmark conventional boost and

benchmark interleaved boost converters and the proposed phase shifted semi-bridgeless boost

converter. The currents in regular diodes Da and Db were assumed to be split with the current

going through intrinsic body diodes for phase shifted semi-bridgeless topology. The regular

diodes input bridge rectifiers have the largest share of losses among the topologies with the input

bridge rectifier. The phase shifted semi-bridgeless topology nearly eliminates this large loss

component (~27.6W). However, the tradeoff is that the MOSFET losses are higher and the

67

intrinsic body diodes of MOSFETs conduct, producing new losses (~7.8W). The fast diodes in

the benchmark conventional and benchmark interleaved PFC have slightly lower power losses,

since the boost RMS current is higher in these topologies.

Figure 3-10: Comparison of the estimated loss distribution in the semiconductors at 70kHz switching

frequency, 240V input, 3300W load at 400V

Overall the MOSFETs are under slightly more stress in phase shifted semi-bridgeless topology,

but the total loss for the proposed phase shifted semi-bridgeless boost are 17% lower than the

benchmark conventional boost and 7% lower than the benchmark interleaved boost . Since the

benchmark converter bridge rectifier losses are large, it is expected that phase shifted semi-

bridgeless boost converter should have the lowest losses among the topologies investigated.

Additionally, it is noted that the losses in the input bridge rectifiers are 63% of total losses in the

conventional PFC converter and 71% of total losses in the benchmark interleaved PFC converter.

27.6

W

12

.9 W

8.3

W

0 W

48.7

W

27.6

W

12

.9 W

8.3

W

0 W

48.7

W

0 W

12

.7 W 1

9.1

W

7.9

W

39.6

W

0

10

20

30

40

50

60

Reg

ula

r

Dio

des

Fa

st D

iod

es

FE

Ts

Intr

insi

c

Bod

y D

iod

es

Tota

l L

oss

es

Pow

er

Loss

es

(W)

Devices / Total Losses

Conventional Boost

Interleaved Boost

Phase Shifted Semi-Bridgeless Boost

68

Therefore, eliminating the input bridge in PFC converters is justified despite that the introduction

of new losses.

3.4. Simulation Results

PSIM simulation software was used to verify steady state waveforms of each component.

Figure 3-11: PSIM simulation circuit for the phase shifted semi-bridgeless PFC boost converter: Vin = 240V,

Vo = 400V, Po = 3400W and fsw = 70 kHz

69

Figure 3-11 shows the PSIM simulation circuits of the proposed BLIL PFC converter. As it can

be seen, the power stage section of converter consists of two boost inductors, Ld1 and Ld2, two

fast boost diodes, Db1 and Db2, two switches, Q1 and Q2 and their body diodes Dq1 and Dq2.

Also it consists of two current loops and one voltage loop. The sensed input voltage is multiplied

by the compensated output voltage, and then generates a control signal to be compared with the

switching carrier waveforms to generate the gating signals for main MOSFETs.

Figure 3-12 shows the PSIM simulation results of a phase shifted semi-bridgeless PFC

boost converter. The input current is in phase with the input voltage, and it has close to unity

power factor. Also the output voltage is regulated at around 400V, with a 120 Hz low frequency

ripple. The converter is operating at 70 kHz switching frequency, 240 V input voltage and 3.4

kW output power.

Figure 3-12: Simulation waveforms for the proposed phase shifted semi-bridgeless PFC boost converter:

Output voltage, input voltage and input current: Vin = 240V, Vo = 400V, Po = 3400W and fsw = 70 kHz

70

3.5. Logic Implementation

A standard two-phase interleaved CCM PFC controller from Texas Instrument, UCC28070

was used to implement the logic circuitry of the prototype unit.

A simplified logic implementation of this controller applied to semi-bridgeless phase shifted

PFC boost converter is given in Figure 3-13.

Figure 3-13: Logic implementation of phase shifted semi-bridgeless PFC boost converter using UCC28070

controller

71

3.6. Experimental Results

Prototypes of a phase shifted bridgeless boost converter and a benchmark interleaved

boost converter were built to verify the proof-of-concept and analytical work presented in this

chapter and to benchmark the proposed converter. The devices used in experimental prototypes

are provided in Table 3-1.

Table 3-1: Component / devices used in prototype unit

Topology Components Used in Prototype Unit Head

Device Part # / Value # of Devices

Ph

ase

Shif

ted S

emi-

bri

dgel

ess

PF

C Regular Diode 25ETS08S 2

Fast Diode IDB06S60C 2

MOSFET IPB60R099CP 2

Inductors

Kool mu 77071 core

60 turns / 400 µH

2

Capacitors EKXJ451ELL820 / 82 µF 10

Ben

chm

ark I

nte

rlea

ved

PF

C

Regular Diode 25ETS08S 4

Fast Diode IDB06S60C 2

MOSFET IPB60R099CP 2

Inductors

Kool mu 77071 core

60 turns / 400 µH

2

Capacitors EKXJ451ELL820 / 82 µF 10

Pictures of the proposed phase shifted bridgeless boost prototype are provided in Figure 3-14

to Figure 3-16. It consists of a capacitor bank of 820 µF, a control board and an IMS power

board attached to a heatsink with the PFC inductors.

72

Figure 3-14: Capacitor bank of 820 uF

Figure 3-15: Control board of PFC boost converter

73

Figure 3-16: IMS power board attached to a heatsink with the PFC inductors

The experimental efficiency of the phase shifted bridgeless boost converter and

benchmark interleaved boost converter is provided in Figure 3-17 for 240V input and Figure 3-

19 for 120V input at 70 kHz switching frequency and 400 V output. Loss reduction curves as a

function of output power are provided in Figure 3-18 and Figure 3-20 for 240V and 120V input,

respectively.

74

Figure 3-17: Efficiency as a function of output power at Vin = 240V, Vo=400V and 70kHz switching

Figure 3-18: Loss reduction as a function of output power at Vin = 240V, Vo=400V and 70kHz for the

proposed phase shifted semi-bridgeless converter compared to the benchmark interleaved boost converter

94

95

96

97

98

99

100

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

Eff

icie

ncy

(%

)

Output Power (W)

Interleaved PFC Converter

Phase Shifted Semi-Bridgeless

PFC Converter

0

10

20

30

40

50

60

70

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

Loss

Re

du

ctio

n (

%)

Output Power (W)

75

Figure 3-19: Efficiency as a function of output power at Vin = 120V, Vo=400V and 70kHz

Figure 3-20: Loss reduction as a function of output power at Vin = 120V, Vo=400V and 70kHz for the

proposed phase shifted semi-bridgeless converter compared to the benchmark interleaved boost converter

From the results, it is noted that proposed semi-bridgeless PFC converter achieves a peak

efficiency of 98.6% at 1 kW output power. Additionally, the light load efficiency of the

91

92

93

94

95

96

97

98

0

20

0

40

0

60

0

80

0

10

00

12

00

14

00

16

00

18

00

Eff

icie

ncy

(%

)

Output Power (W)

Interleaved PFC Converter

Phase Shifted Semi-Bridgeless

PFC Converter

0

10

20

30

40

50

60

70

0

20

0

40

0

60

0

80

0

10

00

12

00

14

00

16

00

18

00

Loss

Re

du

ctio

n (

%)

Output Power (W)

76

proposed converter is significantly better than that of the benchmark interleaved PFC due to the

absence of the input bridge rectifier. However, as the load increases, the efficiency drops due to

additional heat dissipation in the intrinsic body diodes of the MOSFETs.

Figure 3-21: THD as a function of output power at Vin = 120 V and 240V, Vo=400V and 70kHz

Figure 3-22: Power factor as a function of output power at Vin = 120 V and 240V, Vo=400V and 70kHz

0

5

10

15

20

25

30

35

40

45

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

TH

D (

%)

Output Power (W)

Vin=240

Vin=120

0.84

0.86

0.88

0.9

0.92

0.94

0.96

0.98

1

1.02

0

50

0

10

00

15

00

20

00

25

00

30

00

35

00

Po

we

r fa

cto

r

Output Power (W)

Vin=240

Vin=120

77

Figure 3-23: Harmonics orders at Vin = 120 V and 240V, compared against EN61000-3-2 standard for Vin =

120 V at Po = 1700 W and 240 V at Po = 3400 W

In order to verify the quality of the input current, the input current THD is shown in Figure 3-21.

The power factor and harmonic orders are given and compared with EN 61000-3-2 standard in

Figure 3-22 and Figure 3-23. It is noted that mains current THD is less than 5% from 50% load

to full load and it is compliant to IEC 61000-3-2 (Figure 3-21 and Figure 3-23). The converter

power factor is shown over entire load range for 120 and 240V input in Figure 3-22. The power

factor is greater than 0.99 from 50% load to full load.

Experimental waveforms from the proposed converter prototype are provided in Figure 3-24

through Figure 3-30. The input current, input voltage and output voltage are given in Figure 3-

24. As it can be seen, the input current is in phase with the input voltage and has a sinusoidal

shape. Additionally, there is a low frequency ripple on output voltage, which is inversely

proportional to the value of PFC bus output capacitors.

0

0.5

1

1.5

2

2.5

3 5 7 9

11

13

15

17

19

21

23

25

27

29

31

33

35

37

39

Am

pli

tud

e (

A)

Harmonics Order

EN 61000-3-2 Class D Limits (A)

Amplitude (A) Vin = 120 V

Amplitude (A) Vin = 240 V

78

In Figure 3-25, the inductor current is provided in addition to the above mentioned waveforms

from Figure 3-24. It is noted that during the positive half-cycle, the inductor current is the same

as the input current. However, during the negative half-cycle, the input current is partially

flowing through slow diodes, Da and Db.

In Figure 3-26 the inductor current, input current and current sensed in the MOSFET through a

current transformer are given. The gating signals, sensed MOSFET current and the inductor

current are provided for duty cycles less than 0.5, Figure 3-27, and greater than 0.5, Figure 3- 28.

These waveforms match the theoretical models.

Figure 3-29 and Figure 3-30 show the output voltage transient response to a change to the load

form FL to NL and vice versa. As can be noted, the output voltage regulates right away.

Figure 3-24: Input current, input voltage and output voltage.

Input Voltage

Ch2 = Vin 100V/div.

Output Voltage Ch1= Vo 100V/div.

Input Current

Ch4 = Iin 10A/div.

79

Figure 3-25: Input current, inductor current, input voltage and output voltage.

Figure 3-26: Inductor current, input current and sensed MOSFET current.

Input Voltage

Ch2 = Vin 100V/div.

Output Voltage Ch1= Vo 100V/div.

Inductor Current

Ch3 = IL1 10A/div.

Input Current

Ch4 = Iin 10A/div.

Sensed MOSFET Current Ch1= IQ1 2V/div.

Inductor Current

Ch3 = IL1/IDb 10A/div.

Input Current

Ch4 = Iin 10A/div.

80

Figure 3-27: Gating signal, inductor and sensed MOSFET current for D < 0.5

Figure 3-28: Gating signal, inductor and sensed MOSFET current for D > 0.5

Sensed MOSFET Current Ch2= IQ1 2V/div.

Inductor Current

Ch3 = IL1/IDb 10A/div.

Gating Signal

Ch1= Vg 10V/div.

Sensed MOSFET Current Ch2= IQ1 2V/div.

Inductor Current

Ch3 = IL1/IDb 10A/div.

Gating Signal

Ch1= Vg 10V/div.

81

Figure 3-29: Load transient response from NL to FL (0W to 3400 W)

Ch1= Vo 100V/div. Ch2= Iin 10A/div. Ch4= Vo 100V/div

Figure 3-30: Load transient response from FL to NL (3400 W to 0 W)

Ch1= Vo 100V/div. Ch2= Iin 10A/div. Ch4= Vo 100V/div

82

3.7. Conclusion

A new high performance phase shifted semi-bridgeless AC-DC Boost converter topology

has been presented in this chapter for the front-end AC-DC converter in PHEV battery chargers.

The proposed converter features high efficiency at light loads and low lines, which is critical to

minimize the charger size, charging time and the amount and cost of electricity drawn from the

utility; the component count, which reduces the charger cost; and reduced EMI. The converter is

ideally suited for automotive level 1 residential charging applications in North America where

the typical supply is limited to 120 V and 1.44 kVA.

An analysis and performance characteristics are presented. A breadboard converter circuit

has been built to verify the proof-of-concept. The theoretical waveforms were compared with the

results taken from prototype unit. Additionally, key experimental waveforms were provided and

input current harmonics at each harmonic order were compared more explicitly with the IEC

61000-3-2 standard limits.

Experimental results demonstrate that the mains current THD is smaller than 5% from

50% load to full load and the converter is compliant with the IEC 61000-3-2 standard. The

converter power factor was also provided for full power range at 120 and 240 V input. The

power factor is greater than 0.99 from 50% load to full load. The proposed converter achieves a

peak efficiency of 98.6 % at 240 V input and 1 kW output power.

83

CHAPTER 4. The Ripple Steering Technique and Converter

Modeling

4.1. Introduction

In this chapter, an average switch model approach to the power stage modeling, feedback

compensation, and dynamic analysis of PFC boost converters with coupled magnetic filter is

presented. The model is expressed by the derivation of power stage transfer functions for

conventional boost converter, and then followed by the power stage transfer functions for PFC

boost converter with coupled magnetic filters. Experimental and simulation results of a prototype

boost converter converting universal AC input voltage to 400 V DC at 1.8 kW are given to verify

the proof of concept, and analytical work reported in this chapter. The experimental result

demonstrates that the model can correctly predict the steady-state and large signal dynamic

behavior of a CCM PFC boost converter with coupled magnetic filter.

4.2. Average Switch Model

Using the model discussed in section 1.3.6 applying that to modified boost converter with

coupled inductors, any of the transfer functions between the output variables (output voltage and

inductor current) and the input variables (input voltage and duty ration) can be derived.

4.2.1. PWM Switch Model of Conventional Boost Converter

Applying the PWM switch model results in the equivalent circuit shown in Figure 4-1 for the

conventional boost converter:

84

Figure 4-1: Conventional boost converter with PWM switch

By application of superposition and other circuit theory, the power stage transfer functions can

be derived:

;A$ >?.h .[= :-&h .Ai.[= :+-Ai :-h&,0+ 4-1

5$ >?.h&,0h .Ai.[= :+-Ai :-h&,0+ 4-2

The low voltage loop cut off frequency of around 10 Hz is well below the LC resonant frequency

(around 125 Hz), so the voltage loop transfer function can be approximated by:

5$ >?ɳ.∆kl.h .[= : 4-3

4.2.1. Feedback Compensation Design of Conventional Boost Converter

The compensation is selected so that the open-loop transfer function verifies the following

criteria:

1- A high gain at low frequency in order to compensate the steady state error.

2- A gain slope maintained at -20 dB/dec around the crossover frequency in order to ensure

enough phase margin and, therefore, closed-loop stability.

3- A very small gain at high frequency in order to reduce the influence of the switching

harmonics and overall noise.

85

A. Current Loop Compensation

Figure 4-2 shows the current loop plant and a type II compensator Bode plots. The current loop

power stage has a cross over frequency at 10 KHz, so a type II compensator was chosen with the

following pole and zero, as shown in Figure 4-3:

Figure 4-2: Current loop plant and compensator type II Bode plots

LB = 400 µH, Co = 1000 µF, RL = 88 Ω, R2 = 11.5 kΩ, C2 = 4.7 nF and C3 = 230 pF.

C3

C2R2

R1

Figure 4-3: Type II compensator network

10 100 1 103

× 1 104

× 1 105

× 1 106

× 1 107

×

75−

50−

25−

0

25

50

75

180−

150−

120−

90−

60−

30−

0

Frequency (Hz)

Mag

nit

ude

(dB

)

Phas

e (D

egre

e)

Plant Transfer Function

Phase and Magnitude

Controller Type II

Phase and Magnitude

86

The compensator has a pole at zero frequency, and another pole at m &HP+[+ nEn+o nE and one

zero at mp &HP+[+ .

The values for compensator are: R2 = 11.5 KΩ, C2 = 4.7 nF and C3 = 230 pF.

The open loop bode plot is given in Figure 4-4. As it can be seen, at cross over frequency, the

open loop has a phase margin of 65° which is a stable design.

Figure 4-4: Open loop Bode plot for current loop

LB = 400 µH, Co = 1000 µF, RL = 88 Ω, R2 = 11.5 kΩ, C2 = 4.7 nF and C3 = 230 pF.

B. Voltage Loop Compensation

The voltage loop transfer function and compensator network Bode plot is given in Figure 4-5.

The compensator for the voltage is a type II compensator as well. In the compensator network,

voltage error amplifier gain is adjusted with compensation components to attenuate the twice-

1 103

× 1 104

× 1 105

× 1 106

×

100−

80−

60−

40−

20−

0

20

40

60

80

100

180−

162−

144−

126−

108−

90−

Frequency (Hz)

Mag

nit

ude

(dB

)

Phas

e (D

egre

e)

Compensated Current Loop

Plant (Open Loop)

Phase and Magnitude

87

line frequency ripple on the output capacitor to obtain a desired reduction of 3rd harmonic THD.

C3 sets the reduction level, R2 sets the phase margin to 45 degrees at cross over frequency, and

C2 sets the beginning of the phase boost. The desired cross over frequency for voltage loop is

around 10 Hz, where the line frequency varies from 50 Hz to 60 Hz.

The values for compensator are: R2 = 75 KΩ, C2 = 1 µF and C3 = 100 nF.

Figure 4-6 shows the open loop Bode plot of compensated voltage loop. As it can be seen, it has

a high gain at low frequencies, and at cross over frequency, the open loop has a phase margin of

55° which is a stable design.

Figure 4-5: Voltage loop plant and compensator type II Bode plots

LB = 400 µH, Co = 1000 µF, RL = 88 Ω, R2 = 75 kΩ, C2 = 1 µF and C3 = 100 nF.

0.1 1 10 100 1 103

×

100−

80−

60−

40−

20−

0

20

40

60

80

100

180−

144−

108−

72−

36−

0

Frequency (Hz)

Mag

nit

ud

e (d

B)

Ph

ase

(Deg

ree)Plant Transfer Function

Phase and Magnitude

Controller Type II

Phase and Magnitude

88

Figure 4-6: Open loop Bode plot for voltage loop

LB = 400 µH, Co = 1000 µF, RL = 88 Ω, R2 = 75 kΩ, C2 = 1 µF and C3 = 100 nF.

4.3. Ripple Steering Techniques in PFC Application

Although the control strategy of PFC stage is similar to that of a conventional boost

converter, but their power stage transfer functions are different. And no modeling has been done

to verify the effect of added coupled filter to the power stage transfer functions, and thus the

design of feedback loop compensator.

Figure 4-7: Modified PFC boost converter with coupled inductors

0.1 1 10 100 1 103

×

100−

80−

60−

40−

20−

0

20

40

60

80

100

180−

153−

126−

99−

72−

45−

Frequency (Hz)

Mag

nit

ud

e (d

B)

Ph

ase

(Deg

ree)

Compensated Voltage Loop

Plant (Open Loop)

Phase and Magnitude

89

Application of PWM switch model discussed in section 1.3.6 is used for the circuit

shown in Figure 4-7 for boost converter with coupled inductors in order to get the voltage and

current plant transfer functions.

4.3.1. PWM Switch Model of Boost Converter with Coupled Inductor

In order to model the boost converter with coupled inductors, the winding arrangements

shown in Figure 4-8 can be described mathematically by:

5& q& 6 + r 6+ 4-4

5H qH 6+ + r 6 4-5

where: r s tq& qH 4-6

M is the mutual inductance and k is the coupling coefficient of the windings.

Figure 4-8: Equivalent circuit of coupled inductors

From above equations, Figure 4-8 can be replaced with its equivalent circuit shown in Figure 4-9

[71].

Figure 4-9: Equivalent circuit of coupled inductors

90

The equivalent circuit of coupled inductor can be shown according to:

5& q& − r 6 + r6 + 6+ 4-7

5H qH − r 6+ + r6 + 6+ 4-8

Figure 4-10: Modified boost converter with PWM switch

Figure 4-10 shows the equivalent circuit of boost converter with a coupled inductor,

where active and passive switches are replaced by the averaged model of PWM switch. By

application of superposition and other circuit theory, the power stage transfer functions can be

derived:

;A$ >?.h .[= :-&.[[=A+,v :+-&]0: 4-9

5$ >?.h&,0.[[=A+,v :+-&]0: 4-10

Where:

)<$ [xA. y . y:. rqH − 3r] $Z + [y:. r2qH − 3r + xA . y . qH] $D + xA[y:. qH1 −)H + 2y:. r1 − )H + 2y . r] $H + 2r $ + xA1 − )H 4-11

The voltage loop transfer function approximation is valid here as well, so equation (4-5) can

be simplified as equation (4-3). Figure 4-11 shows the Bode plot of current loop transfer

functions for modified boost converter with coupled inductor and a conventional boost converter.

91

As it can be noted, the phase margins for both plants are the same and steady at -90 °. But the

main difference is the magnitude of current plant transfer functions, where for conventional

boost has lower cross over frequency and lower gain. If the same feedback compensation

network is used for both converters, the steady state and large signal responses would not be

adequate.

Figure 4-11: Current loop plant Bode plots

LB = 400 µH, Co = 1000 µF, RL = 88 Ω, Ldc = 400 µH, Lac = 250 µH, M = 260 µH, Cs = 1 µF.

1 103

× 1 104

× 1 105

× 1 106

× 1 107

×

75−

50−

25−

0

25

50

75

180−

150−

120−

90−

60−

30−

0

Frequency (Hz)

Mag

nit

ud

e (d

B)

Ph

ase

(Deg

ree)

Conventional Boost

Plant Transfer Function

Phase and Magnitude

Modified Boost Plant

Transfer Function

Phase and Magnitude

92

4.3.1. Feedback Compensation Design of Boost Converter with Coupled

Inductor

Figure 4-12 shows the current loop plant and a type II compensator Bode plots. The

current loop power stage has a cross over frequency at 20 KHz, so a type II compensator was

chosen with the following pole and zero: pole at m &HP+[+ nEn+o nE and one zero at mp &HP+[+ .

The values for compensator are: R2 = 22.1 KΩ, C2 = 1 nF and C3 = 150 pF.

The open loop Bode plot is given in Figure 4-13. As it can be seen, at the cross over frequency,

the open loop has a phase margin of 50° which is a stable design.

Figure 4-12: Current loop plant and compensator Bode plots for boos converter with coupled inductor

RL = 88 Ω, Ldc = 400 µH, Lac = 250 µH, M = 260 µH, Cs = 1 µF, R2 = 22.1 kΩ, C2 = 1 nF and C3 = 150 pF.

10 100 1 103

× 1 104

× 1 105

× 1 106

× 1 107

×

50−

25−

0

25

50

75

100

180−

150−

120−

90−

60−

30−

0

Frequency (Hz)

Mag

nit

ud

e (d

B)

Ph

ase

(Deg

ree)

Modified boost Plant

Transfer Function

Phase and Magnitude

Controller Type II

Phase and Magnitude

93

Figure 4-13: Open loop plant Bode plot for boost converter with coupled inductor

RL = 88 Ω, Ldc = 400 µH, Lac = 250 µH, M = 260 µH, Cs = 1 µF, R2 = 22.1 kΩ, C2 = 1 nF and C3 = 150 pF.

4.4. Simulation Results

PSIM simulation was used to verify the feedback loop design for steady state and large signal

perturbation. Figure 4-14 shows the simulation results of boost converter with coupled inductors,

converting universal AC input voltage to 400 V DC at 240 V input and 1.6 kW. Figure 4-15

illustrates: Top: The boost inductor current in a conventional boost converter - No Filtering technique.

Middle: The boost inductor current with ripple steering technique applied, and Bottom: Series capacitor

current in the coupled inductor filter.

10 100 1 103

× 1 104

× 1 105

× 1 106

×

100−

75−

50−

25−

0

25

50

75

100

125

150

200−

180−

160−

140−

120−

100−

Frequency (Hz)

Mag

nit

ude

(dB

)

Phas

e (D

egre

e)

Compensated Current Loop

Plant (Open Loop)

Phase and Magnitude

94

Figure 4-14: PSIM simulation circuit for ripple steering technique applied to PFC boost converter

Figure 4-15: Inductor Current - no filtering technique (Top) – ripple steering technique (Middle), series

capacitor current (Bottom): Vin = 240 V, Vo = 400 V, Po = 1600 W, fsw = 70 kHz

0

-2

2

4

6

8

10

12

0

-2

2

4

6

8

10

12

I(M1_1)

0.16 0.17 0.18 0.19 0.2

0

-2

-4

2

4

I(M1_2)

95

4.5. Experimental Results

Prototypes of the boost converter with coupled inductors and conventional boost converter

were built to verify the proof-of-concept and analytical work presented in this thesis and to

benchmark the proposed compensation network design. Figure 4-16 shows a coupled inductor

used for this experiment. The inductor values are: Ldc = 400 uH, Lac = 250 uH, M = 265 uH and

the series capacitor Cs = 1 uF. Figure 4-17 and Figure 4-18 show the experimental and

simulation ripple current waveform in dc inductor under the following operating conditions: Vin

= 120 V, Iin = 15 A, Po = 800 W, Vo = 400 V and 70 kHz.

Figure 4-16: Coupled inductors used in experimental circuit - Top left: DC inductor, Top right: AC inductor,

Bottom: Coupled inductor

96

Figure 4-17: Inductor current Idc ripple at 120 V input and 800 W output - Experimental

Figure 4-18: Inductor current Idc ripple at 120 V input and 800 W output - Simulation

97

Figure 4-19: Inductor current Iac at 120 V input and 800 W output - Experimental

Figure 4-20: Inductor current Iac at 120 V input and 800 W output - Simulation

0.17 0.175 0.18 0.185 0.19 0.195 0.2

Time (s)

0

-2

-4

-6

-8

2

4

6

8ac Inductor Current

98

Figure 4-21: Peak inductor current Iac ripple at 120 V input and 800 W output - Experimental

Figure 4-22: Peak inductor current Iac ripple at 120 V input and 800 W output - Simulation

99

Figure 4-23: Inductor current Idc ripple at 240 V input and 1600 W output - Experimental

Figure 4-24: Inductor current Idc ripple at 240 V input and 1600 W output - Simulation

100

Figure 4-25: Inductor current Iac at 240 V input and 1600 W output - Experimental

Figure 4-26: Inductor current Iac at 240 V input and 1600 W output - Simulation

101

Figure 4-27: Peak inductor current Iac ripple at 240 V input and 1600 W output - Experimental

Figure 4-28: Peak inductor current Iac ripple at 240 V input and 1600 W output - Simulation

102

Figure 4-19 and Figure 4-20 show the experimental and simulation current waveforms in ac

inductor under the following operating conditions: Vin = 120 V, Iin = 15 A, Po = 800 W, Vo =

400 V and 70 kHz. Figure 4-21 and Figure 4-22 show the experimental and simulation ripple

current waveforms in ac inductor under the same operating conditions as above.

Figure 23 through Figure 28 repeats the same waveforms for different test condition: Vin =

240 V, Iin = 15 A, Po = 1600 W, Vo = 400 V and 70 kHz.

4.6. Conclusion

An averaged PWM model for boost converters with ripple steering technique has been

developed and effect of ripple steering technique on properly design of compensation network

for the current loop has been studied. The analytical and simulation results were compared with

experimental results of a breadboard converter circuit converting universal AC input voltage to

400 V DC at 240 V input and 1.6 kW. The proposed model shows an accurate prediction of

steady state and large transient behavior of boost converter with coupled inductors.

103

CHAPTER 5. Conclusions and Future Work

5.1. Conclusions

As the adaptation rate of PHEVs increases, the stress on the utility grid is projected to

increase significantly at times of peak demand. Therefore, efficient and high power factor

charging is critical in order to minimize the utility load stress, and reduce the charging time. In

addition, a high power factor is needed to limit the input current harmonics drawn by these

chargers and to meet regulatory standards.

To meet the next generation requirements of these applications, two new topologies and

two new modeling techniques have been proposed in this thesis.

5.1.1. Bridgeless Interleaved Boost PFC Converter

The first contribution is a new bridgeless interleaved boots PFC converter for plug in

hybrid electric vehicle level 2 charger applications. The proposed topology achieves a peak

efficiency of 98.94% was reached at 265 V input and 1.2 kW output power. High efficiency over

the entire load range is achieved in this topology, enabling fewer problems with heat dissipation

and cooling systems. Furthermore, a higher efficiency means more power is available from the

mains feed to charge the batteries, reducing charging times and electricity costs.

5.1.2. Phase Shifted Semi-Bridgeless Boost PFC Converter

The second contribution is a new phase shifted semi-bridgeless boost PFC converter for

plug in hybrid electric vehicle level 1 charger applications. Experimental results demonstrate that

the mains current THD is smaller than 5% from 50% load to full load and the converter is

compliant with the IEC 61000-3-2 standard. The converter power factor was also provided for

104

full power range at 120 and 240V input. The power factor is greater than 0.99 from 50% load to

full load. The proposed converter achieves a peak efficiency of 98.6 % at 240 V input and 1 kW

output power.

5.1.3. New Loss Modeling for PFC Boost Converters

In order to properly select the power stage components of a converter and calculate the

associated power losses, it is necessary to determine the RMS and average values of their

currents. In a typical boost converter, the MOSFET and diode current waveforms are pulsed-

width modulated, with both the duty cycle and peak amplitude varying with the AC input. And

without an effective mathematical method for computing these RMS and average values, the

proper design and selection of power stage components can be flawed. A new analytical model

for the four different topologies was developed, enabling the calculation of power losses and

efficiency calculation.

5.1.1. New Average Modeling for PFC Boost Converters with Coupled

Inductors

An averaged PWM model for boost converters with ripple steering technique has been

developed and effect of ripple steering technique on properly design of compensation network

for current loop has been studied. The analytical and simulation results were compared with

experimental results of a breadboard converter circuit converting universal AC input voltage to

400 V DC at 240 V input and 1.8 kW. The proposed model shows an accurate prediction of the

steady state and large transient behavior of boost converter with coupled inductor.

105

5.2. Future Work

This sub-section outlines the possible future work for the thesis topics.

5.2.1. LLC Resonant Converter for DC/DC Stage

The front end PFC section of a PHEV charger is followed by a DC/DC section to

complete the charger system. This stage can be an LLC resonant converter. High efficiency (>

98%), high output voltage (~ 450 V), high power (~ 3.3 KW) and wide output voltage range (220

V to 450 V) makes it a challenging research topic.

5.2.2. Resonant PFC Converter

Most of hard switching PFC boost topologies have been discussed in this dissertation, but

a new class of topology for PFC is resonant PFC converters. Recent research activities have been

conducted by Infineon Technology on Bridgeless Interleaved resonant PFC and also by Dr.

Slobodan Ćuk from TESLAco on Bridgeless Resonant PFC.

5.2.3. Level 3 Chargers

With cities having it mandatory to have a Level 3 charger installed in every high rise build

in the lower main island in a near future, the business case is already justified.

The ultimate in charge technology, but nobody has yet really settled on a standard, leaving

manufacturers of cars and charging equipment to come up with their own. This type of charging

is defined as any charging above 14.4 kilowatts. Also, most Level 3 stations are considered “fast

chargers,” but they don’t have to be. The conventional wisdom is that if an average EV battery

can be charged to full in about a half hour, then it is a fast charge.

Interleaved resonant PFC and also by Dr. Slobodan Ćuk from TESLAco on Bridgeless

106

5.2.4. Wireless Chargers

Wireless charger is another interesting topic for future research. Several companies are

investing to develop a family of wireless electric power components that will enable OEM’s in a

broad range of industries and applications to make their products truly “wireless.” Wireless

electric power delivered over room scale distances, and with high efficiency. Wireless electric

power that is safe for people and animals. This might be the future of EV and PHEV chargers.

107

Bibliography

[1] K. Morrow ; D. Karner ; J. Francfort, "Plug-in Hybrid Electric Vehicle Charging Infrastructure

Review," U.S. Departent of Energy - Vehicle Technologies Program, 2008.

[2] "Compliance testing to the IEC 1000-3-2 (EN 61000-3-2) and IEC 1000-3-3 (EN 61000-3-3)

Standards ": Agilent Technology.

[3] A. F. Lisheng Shi; Meintz, M.; , "Single-Phase Bidirectional AC-DC Converters for Plug-in Hybrid

Electric Vehicle Applications," IEEE Vehicle Power and Propulsion Conference, VPPC, pp. 1 - 5

2008.

[4] http://www.eere.energy.gov/, "US Department of Energy - The Office of Energy Efficiency and

Renewable Energy (EERE)."

[5] B. Singh ; B.N. Singh ; A. Chandra ; K. Al-Haddad ; A. Pandey ; D.P. Kothari, "A Review of Single-

Phase Improved Power Quality AC-DC Converters," IEEE Transactions on Industrial Electronics,

vol. 50, pp. 962 - 981, 2003.

[6] C. Qiao ; K.M. Smedley, "A Topology Survey of Single-Stage Power Factor Corrector with a Boost

Type Input-Current-Shaper," IEEE Transactions on Power Electronics, vol. 16, pp. 360 - 368, 2001.

[7] D. Tollik ; A. Pietkiewicz, "Comparative analysis of 1-phase active power factor correction

topologies," in IEEE Telecommunications Energy Conference, INTELEC, 1992, pp. 517 - 523

[8] L. Petersen ; M. Andersen, "Two-Stage Power Factor Corrected Power Supplies: The Low

Component-Stress Approach " in IEEE Applied Power Electronics Conference and Exposition,

APEC. vol. 2, 2002, pp. 1195 - 1201.

[9] T.S. Key ; Jih-Sheng Lai, "IEEE and International Harmonic Standards Impact on Power Electronic

Equipment Design," in International Conference on Industrial Electronics, Control and

Instrumentation, IECON. vol. 2, 1997, pp. 430 - 436.

[10] Ned Mohan ; Tore M. Undeland, Power electronics: converters, applications, and design: Wiley

India, 2009.

[11] "IEEE Std 519-1992 IEEE Recommended Practices and Requirements for Harmonic Control in

Electrical Power Systems," IEEE 1992.

[12] "Guide to Energy Management - Power Factor," BC Hydro, Vancouver 2000.

[13] L. Huber ; Liu Gang ; M.M. Jovanovic, "Design-Oriented Analysis and Performance Evaluation of

Buck PFC Front End," vol. 25, pp. 85 - 94, 2010.

[14] A. Fernandez ; J. Sebastian ; M.M. Hernando ; P. Villegas ; J. Garcia, "Helpful Hints to Select a

Power-Factor-Correction Solution for Low- and Medium-Power Single-Phase Power Supplies,"

IEEE Transactions on Industrial Electronics, vol. 52, pp. 46 - 55 2005.

[15] J. M. Bourgeois, "CIRCUITS FOR POWER FACTOR CORRECTION WITH REGARDS TO MAINS

FILTERING," STMicroelectronics 1999.

[16] R. Randall, "Choosing Power Switching Devices for SMPS Designs MOSFETs or IGBTs?," Fairchild

Semiconductor Application Note AN-7010 2005.

[17] L. Rossetto ; G. Spiazzi ; P. Tenti, "CONTROL TECHNIQUES FOR POWER FACTOR CORRECTION

CONVERTERS."

[18] W. Z. G. F. Y.-F. L. B. Wu;, "A Digital Power Factor Correction (PFC) Control Strategy Optimized

for DSP," IEEE Transactions on Power Electronics vol. 19, pp. 1474 - 1485, 2004.

[19] P. N. Sangsun Kim; Enjeti, "Control of Multiple Single-Phase PFC Modules With a Single Low-Cost

DSP," IEEE Transactions on Industry Applications, vol. 39, pp. 1379 - 1385 2003.

[20] M. C. J. Sun;, "Feedforward current control of boost single-phase PFC converters," IEEE

Transactions onPower Electronics, vol. 21, pp. 338 - 345, 2006.

108

[21] A. J. S. Min Chen; Mathew, "Nonlinear Current Control of Single-Phase PFC Converters," Power

Electronics, IEEE Transactions on vol. 22, pp. 2187 - 2194, 2007.

[22] Y. Chen ; D. He ; R.M. Nelms, "Control of a Single-Phase PFC Preregultaor using an 8-bit

Microcontroller," in IEEE Applied Power Electronics Conference, APEC, 2007, pp. 1454 - 1460

[23] P. Grasblum, "Implementing PFC Average Current Mode Control using the MC9S12E128,"

Freescale Semiconductor Application Note AN3052 2005.

[24] "Implementing a Digital AC/DC Switched-Mode Power Supply using a 56F8300 Digital Signal

Controller," Freescale Semiconductor Application Note AN3115 2005.

[25] G. Chu ; S.C. Tan ; C.K. Tse ; S.C. Wong, "General Control for Boost PFC Converter from a Sliding

Mode Viewpoint," in IEEE Power Electronics Specialists Conference, PESC, 2008, pp. 4452 - 4456.

[26] Y. Z. W. X. Y. Yu;, "The PFC with Average Current-Mode and Voltage Fuzzy Controller for the

Output Voltage," in International Symposium on Intelligent Information Technology

Application,IEEE IITA, 2008, pp. 771 - 775.

[27] J.M. Carrasco ; J.M. Quero ; F.P. Ridao ; M.A. Perales ; L.G. Franquelo, "Sliding Mode Control of a

DC-DC PWM Converter with PFC Implemented by Neural Netwroks," IEEE Transactions on

Circuits and Systems I: Fundamental Theory and Applications, vol. 44, pp. 743 - 749 1997.

[28] K. M. Zheren Lai; Smedley, "A Family of Continuous-Conduction-Mode Power-Factor-Correction

Controllers Based on the General Pulse-Width Modulator," IEEE Transactions on Power

Electronics, vol. 13, pp. 501 - 510 1998

[29] S. Basu ; T.M. Undeland, "Inductor Design Considerations for optimizing performance & cost of

Continuous Mode Boost PFC Converters," in IEEE Applied Power Electronics Conference and

Exposition, APEC. vol. 2, 2005, pp. 1133 - 1138.

[30] L. Dixon, "Average Current Mode Control of Switching Power Supplies," UNITRODE

CORPORATION Application Note SLUP091 1990.

[31] E. Najafi ; A. Vahedi ; A. Mahanfar ; A. Yatim, "A new controlling method based on peak current

mode (PCM) for PFC," in IEEE Power and Energy Conference, PECon, 2008, pp. 1103 - 1107

[32] C. Zhou ; R.B. Ridley ; F.C. Lee, "Design and Analysis of a Hysteretic Boost Power Factor

Correction Circuit," in IEEE Power Electronics Specialists Conference, PESC, 1990, pp. 800 - 807.

[33] J. P. Noon, "Designing High-Power Factor Off-Line Power Supplies," Texas Instruments

Application Note SLUP203 2003.

[34] P. C. Todd, "UC3854 Controlled Power Factor Correction Circuit Design," Unitrode APPLICATION

NOTE SLUA 144 1999.

[35] E.M Hertz ; S. Busquets-Monge ; D. Boroyevich ; M. Arpilliere ; H. Boutillier, "Analysis of the

Trade-offs between Thermal Behavior and EMI Noise Levels in a Boost PFC Circuit," in IEEE

Industry Applications Conference, IAS. vol. 4, 2001, pp. 2460 - 2465.

[36] F. C. O. Shuo Wang; Lee, W.G.; , "Characterization, Evaluation, Design of Noise Separator for

Conducted EMI Noise Diagnosis," IEEE Transactions on Power Electronics vol. 20, pp. 974 - 982

2005.

[37] F. C. Shuo Wang; Pengju Kong; Lee, "Common Mode Noise Reduction for Boost Converters Using

General Balance Technique," IEEE Transactions on Power Electronics, vol. 22, pp. 1410 - 1416

2007.

[38] M. T. L. Wei Zhang; Zhang, F.C.; Roudet, J.; Clavel, E.; , "Conducted EMI Analysis of a Boost PFC

Circuit," in IEEE Applied Power Electronics Conference and Exposition, APEC. vol. 1, 1997, pp. 223

- 229.

[39] L. Rossetto ; S. Buso ; G. Spiazzi, "Conducted EMI issues in a 600-W single-phase boost PFC

design," IEEE Transactions on Industry Applications, vol. 36, pp. 578 - 585 2000.

[40] F. Beltrame ; L. Roggia ; A. da Rosa Abaide ; L. Schuch ; J.R. Pinheiro, "EMI Investigation Yield by

Single-Phase PFC Pre-Regulators," in IEEE Industrial Electronics, IECON, 2009, pp. 4128 - 4133

109

[41] B. Lu ; W. Dong ; S. Wang ; F.C. Lee, "High frequency investigation of single-switch CCM power

factor correction converter," in IEEE Applied Power Electronics Conference and Exposition, APEC.

vol. 3, 2004, pp. 1481 - 1487.

[42] F. C. Pengju Kong; Shuo Wang; Lee, "Improving Balance Technique for High Frequency Common

Mode Noise Reduction in Boost PFC Converters," in IEEE Power Electronics Specialists

Conference, PESC, 2008, pp. 2941 - 2947

[43] F. C. O. Shuo Wang; Lee, W.G.; , "Improving the Performance of Boost PFC EMI Filter," in IEEE

Applied Power Electronics Conference and Exposition, APEC. vol. 1, 2003, pp. 368 - 374.

[44] F. C. O. Liyu Yang; Bing Lu; Wei Dong; Zhiguo Lu; Ming Xu; Lee, W.G.; , "Modeling and

Characterization of a 1KW CCM PFC Converter for Conducted EMI Prediction," in IEEE Applied

Power Electronics Conference and Exposition, APEC. vol. 2, 2004, pp. 763 - 769.

[45] C. Wang ; M. Xu ; B. Lu ; F.C. Lee, "New Architecture for MHz Switching Frequency PFC," in IEEE

Applied Power Electronics Conference, APEC, 2007, pp. 179 - 185

[46] F.C. Lee ; Shuo Wang ; Pengju Kong ; Chuanyun Wang ; Dianbo Fu, "Power architecture design

with improved system efficiency, EMI and power density," in IEEE Power Electronics Specialists

Conference, PESC, 2008, pp. 4131 - 4137.

[47] F. C. Dehong Xu; Jindong Zhang; Weiyun Chen; Jinjun Lin; Lee, "Evaluation of output filter

capacitor current ripples in single phase PFC converters " in Proceedings of the Power

Conversion Conference, PCC. vol. 3 Osaka, Japan, 2002, pp. 1226 - 1231.

[48] D. Mitchell, "AC–DC converter having an improved power factor," United States, 1983.

[49] W.-Y. C. J.-M. K. E.-H. K. J.-J. L. B.-H. Kwon;, "Bridgeless Boost Rectifier With Low Conduction

Losses and Reduced Diode Reverse-Recovery Problems," IEEE Transactions on Industrial

Electronics, vol. 54, pp. 769 – 780, April 2007 2007.

[50] M. M. D. Yungtaek Jang; Jovanovic, D.L.;, "Bridgeless PFC boost rectifier with optimized

magnetic utilization," in IEEE Applied Power Electronics Conference and Exposition, 2008, pp.

1017 – 1021.

[51] J. M. Hancock, "Bridgeless PFC Boosts Low-Line Efficiency," Infineon Technologies, 2008.

[52] U. Moriconi, "A Bridgeless PFC Configuration based on L4981 PFC Controller ":

STMicroelectronics Application Note AN1606, 2002.

[53] B. Lu ; R. Brown ; M. Soldano, "Bridgeless PFC implementation using one cycle control

technique," in IEEE Applied Power Electronics Conference and Exposition. vol. 2, 2005, pp. 812 -

817.

[54] F. C. Pengju Kong; Shuo Wang; Lee, "Common Mode EMI Noise Suppression for Bridgeless PFC

Converters," IEEE Transactions on Power Electronics, vol. 23, pp. 291 – 297, January 2008 2008.

[55] M. M. Yungtaek Jang; Jovanovic, "A Bridgeless PFC Boost Rectifier With Optimized Magnetic

Utilization," IEEE Transactions on Power Electronics, vol. 24, pp. 85 - 93 2009.

[56] C. Petrea ; M. Lucanu, "Bridgeless Power Factor Correction Converter Working at High Load

Variations," in International Symposium on Signals, Circuits and Systems, ISSCS. vol. 2, 2007, pp.

1 - 4

[57] L. Huber ; J. Yungtaek ; M.M. Jovanovic, "Performance Evaluation of Bridgeless PFC Boost

Rectifiers," IEEE Transactions on Power Electronics, vol. 23, pp. 1381 - 1390 2008.

[58] W. R. Frank, M.; Schlenk, M.; , "New Control Methods for Rectifier-Less PFC-Stages," in IEEE

International Symposium on Industrial Electronics, ISIE. vol. 2, 2005, pp. 489 - 493.

[59] M. A. E. T. Qingnan Li; Andersen, O.C.; , "Conduction losses and common mode EMI analysis on

bridgeless power factor correction " in International Conference on Power Electronics and Drive

Systems, PEDS, 2009, pp. 1255 - 1260

[60] P. Kong ; S. Wang ; F.C. Lee, "Common Mode EMI Noise Suppression for Bridgeless PFC

Converters," IEEE Transactions on Power Electronics, vol. 23, pp. 291 – 297, January 2008 2008.

110

[61] M. M. Yungtaek Jang; Jovanovic, "Interleaved Boost Converter With Intrinsic Voltage-Doubler

Characteristic for Universal-Line PFC Front End," IEEE Transactions on Power Electronics, vol. 22,

pp. 1394 – 1401, July 2007 2007.

[62] M. O’Loughlin;, "An Interleaved PFC Preregulator for High-Power Converters." vol. Topic 5: Texas

Instrument Power Supply Design Seminar, 2007, pp. 5-1, 5-14.

[63] L. Balogh ; R. Redl, "Power-factor correction with interleaved boost converters in continuous-

inductor-current mode," in IEEE Applied Power Electronics Conference and Exposition, 1993, pp.

168 - 174

[64] A. Jinsong Zhu; Pratt, "Capacitor ripple current in an interleaved PFC converter," in IEEE Power

Electronics Specialists Conference, 2008, pp. 3444 – 3450.

[65] F. C. Chuanyun Wang; Ming Xu; Lee, "Asymmetrical interleaving strategy for multi-channel PFC,"

in IEEE Applied Power Electronics Conference and Exposition, 2008, pp. 1409 - 1415.

[66] K. Raggl ; T. Nussbaumer ; G. Doerig ; J. Biela ; J.W. Kolar, "Comprehensive Design and

Optimization of a High Power Density Single-Phase Boost PFC," IEEE Transactions on Industrial

Electronics, vol. 56, pp. 2574 - 2587 2009.

[67] M.A.P. Andrade ; L. Schuch ; J.R. Pinheiro, "Generalized Switching Logic Scheme For CCM-PFC

Interleaved Boost Converters," in IEEE Power Electronics Specialists Conference, PESC. vol. 3,

2004, pp. 2353 - 2359.

[68] F. C. C. W. Pengju Kong; Shuo Wang; Lee, "Common-Mode EMI Study and Reduction Technique

for the Interleaved Multichannel PFC Converter," IEEE Transactions on Power Electronics, vol.

23, pp. 2576 - 2584 2008.

[69] F. C. B. L. Chuanyun Wang; Ming Xu; Lee, "EMI Study for the Interleaved Multi-Channel PFC," in

IEEE Power Electronics Specialists Conference, PESC, 2007, pp. 1336 - 1342

[70] D.C. Hamill ;P.T.Krein, "A `zero' ripple technique applicable to any DC converter " in IEEE Power

Electronics Specialists Conference, PESC. vol. 2, 1999, pp. 1165 - 1171.

[71] J.W. Kolar ; H. Sree ; N. Mohan ; F.C. Zach, "Novel aspects of an application of `zero'-ripple

techniques to basic converter topologies " in Power Electronics Specialists Conference, PESC. vol.

1, 1997, pp. 796 - 803.

[72] N.K. Poon ; J.C.P. Liu ; C.K. Tse ; M.H. Pong, "Techniques for input ripple current cancellation:

classification and implementation," IEEE Transactions on Power Electronics, vol. 15, pp. 1144 -

1152 2000.

[73] J. Wang ; W.G. Dunford ; K. Mauch, "Analysis of a ripple-free input-current boost converter with

discontinuous conduction characteristics " IEEE Transactions on Power Electronics, vol. 12, pp.

684 - 694, 1997.

[74] E. Chou ; F. Chen; C. Adragna ; B. Lu, "Ripple steering AC-DC converters to minimize input filter "

in IEEE Energy Conversion Congress and Exposition, ECCE, 2009, pp. 1325 - 1330.

[75] V. Vorperian, "Simplified analysis of PWM converters using model of PWM switch, Part I:

Continuous conduction mode " IEEE Transactions on Aerospace and Electronic Systems, vol. 26,

pp. 490 - 496 1990.

[76] V. Vorperian, "Simplified analysis of PWM converters using model of PWM switch, Part II:

Discontinuous conduction mode," IEEE Transactions on Aerospace and Electronic Systems, vol.

26, pp. 497 - 505 1990.

[77] F. Musavi ; K. Al-Haddad ; H.Y. Kanaan, "A large signal averaged modelling and control of

paralleled DC/DC converters with automatic load sharing," in IEEE Applied Power Electronics

Conference and Exposition, APEC. vol. 2, 2005, pp. 1353 - 1358.

[78] E. Van Dijk ; J.N. Spruijt ; D.M. O'Sullivan ; J.B. Klaassens, "PWM-switch modeling of DC-DC

converters " IEEE Transactions on Power Electronics, vol. 10, pp. 659 - 665 1995.

111

[79] Y.W. Lu ; W. Zhang ; Yanfei Liu, "A large signal dynamic model for single-phase AC-to-DC

converters with power factor correction " IEEE Power Electronics Specialists Conference, PESC,

vol. 2, pp. 1057 - 1063, 2004.

[80] Yan-Fei Liu ; P.C. Sen, "Large-signal modeling of hysteretic current-programmed converters "

IEEE Transactions on Power Electronics, vol. 11, pp. 423 - 430, 1996.

[81] Byungcho Choi ; Sung-Soo Hong ; Hyokil Park, "Modeling and small-signal analysis of controlled

on-time boost power-factor-correction circuit " IEEE Transactions on Industrial Electronics, vol.

48, pp. 136 - 142, 2001.

[82] F. Musavi ; W. Eberle ; W.G. Dunford, "A High-Performance Single-Phase AC-DC Power Factor

Corrected Boost Converter for plug in Hybrid Electric Vehicle Battery Chargers," in IEEE Energy

Conversion Congress and Exposition Atlanta, Georgia, 2010.

[83] F.Musavi ; W. Eberle ; W.G. Dunford, "Efficiency Evaluation of Single-Phase Solutions for AC-DC

PFC Boost Converters for Plug-in-Hybrid Electric Vehicle Battery Chargers," in IEEE Vehicle Power

and Propulsion Conference Lille, France, 2010.

[84] F.Musavi ; W. Eberle ; W.G. Dunford, "A Phase Shifted Semi-Bridgeless Boost Power Factor

Corrected Converter for Plug in Hybrid Electric Vehicle Battery Chargers," in IEEE Applied Power

Electronics Conference and Exposition, APEC Fort Worth, TX, 2011.

112

Appendix

MathCAD Simulation:

Power Component Selection, Magnetic Design and Compensator Network

Design

Part 1 - Power Component Selection

Design Variables Definition

Line Frequency

Minimum Input Voltage

Maximum Input Voltage

High Line Input Voltage

Low Line Input Voltage

Converter Effeciency at High Line

Converter Effeciency at Low Line

Maximum RMS Input Current

Maximum Input Power at High Line

Maximum Input Power at Low Line

Maximum Output Power at High Line

Maximum Output Power at Low Linw

Boost Output Voltage

Minimum Switching Frequency

Maximum boost ripple current as a percentage

Boost FET duty Cycle:

fin 50:=

Vinmin 85:=

Vinmax 265:=

VinHL 230:=

VinLL 115:=

η1 0.985:=

η2 0.965:=

Iin 16:=

Pin1 Iin VinHL⋅:=

Pin2 Iin VinLL⋅:=

Pout1 Pin1 η1⋅:= Pout1 3.625 103

×=

Pout2 Pin2 η2⋅:= Pout2 1.776 103

×=

Vboost 400:=

fs 175 103

⋅:=

∆Irp 50:=

δF_LL θ( ) 12 Vinmin⋅

Vboostsin θ( )⋅−:= δF_HL θ( ) 1

2 Vinmax⋅

Vboostsin θ( )⋅−:=

113

Minimum Duty Cycle at High Line Voltage

Minimum Duty Cycle at Low Line Voltage

Maximum Duty Cycle

Boost Diode Duty Cycle:

PFC Boost Inductor Selection:

Maximum RMS Inductor Current per Phase

Peak Inductor Current per Phase

Boost Inductor Current Ripple

Absolute Peak Inductor Current

0 1.571 3.142 4.712 6.2830

0.2

0.4

0.6

0.8

PFC Boost Converter Duty Ratio v.s. Angle

Angle (Rad.)

Du

ty R

atio

δF_LL θ( )

δF_HL θ( )

θ

DminLL δF_LLπ

2

:= DminLL 0.699=

DminHL δF_HLπ

2

:= DminHL 0.063=

Dmax 0.95:=

δD_LL θ( )2 Vinmin⋅

Vboostsin θ( )⋅:= δD_HL θ( ) 1 δF_HL θ( )−:=

0 1.571 3.142 4.712 6.2830

0.2

0.4

0.6

0.8

Angle (Rad.)

Du

ty R

atio

δD_LL θ( )

δD_HL θ( )

θ

ILIin

28=:=

Ipk 2 IL⋅ 11.314=:=

∆I Lb_LL∆Irp

100Ipk⋅ 5.657=:=

Ipkabs Ipk1

2∆I Lb_LL⋅+ 14.142=:=

114

Current Ripple at High Line

Since Inductor in each Phase will be divided by two equal inductors, so the final indcutor value is 43 uH.

Boost FET Selection:

Instantaneous Boost Inductor Current per Phase

Instantaneous Boost Switch Current

Instantaneous Boost Diode Current

RMS Boost Switch Current

The FET intrinsic body diode current is exactly the same as boost diode current.

RMS Boost Diode Current

Boost FET Loss Calculation:

Conduction Loss

Die Output Capacitor Loss

Gating Losse

Switching Loss

Intrinsic body diode Loss

∆I Lb_HL

Vinmax 2⋅ DminHL⋅

fsLb

2⋅

3.142=:=

iL θ( ) 2 IL⋅ sin θ( )⋅:=

isw θ( ) iL θ( ) δF_LL θ( )⋅:=

iD θ( ) iL θ( ) δD_HL θ( )⋅:=

isw_rms1

2π0

θisw θ( )( )2⌠⌡

d:= isw_rms 5.974=

iD_ave1

2π0

θiD θ( )( )⌠⌡

d

:=iD_ave 0=

Pcond Rds( ) Rds isw_rms2

⋅:=

Pcoss Coss( )1

2Coss⋅ Vboost⋅ fs⋅:=

Pgate Qg Vg, ( ) Qg Vg⋅ fs⋅:=

Psw Tr Tf, ( )1

2isw_rms Tr Tf+( )⋅ Vboost⋅ fs⋅:=

Pid Vid( ) Vid iD_ave⋅:=

115

Boost Inductor Core Selection:

mH - Amperes^2

From the following chart, Kool Mu cores with 60u permability can be used.

Kool Mu Core Selector Chart

Selected CORE 77894

Core Data:

Permability of the core

Window Area cm^2

Path length cm

Volume cm^3

Winding Factor 100% cm/Turn

Lb1

Lb

24.3 10

5−×=:=

Lb1 103

⋅ Ipkabs2

⋅ 8.6= Lb 8.6 105−

×=

µper 60:=

AL 61:=

Aw 1.56:=

le 6.35:=

Ve 4.15:=

MLT 5.23:=

116

Determining Number of Turns:

Assuming 50% roll off, winding for:

Calculated Number of Turns

Final Number of Turns

DC Magnetizing force

From the above curve, the permeability vs DC bias showes a 61% initial permeability at full load.

uH Inductance at FL

uH Inductance at NL

Calculating the Core and Copper Losses:

The wire used is AWG#16:

1

0.5Lb1⋅ 8.6 10

5−×=

Nt

1

0.5Lb1⋅ 1000⋅

AL 106−

37.548=:=

Ntf 25:=

Hor0.4 π⋅ Ntf⋅ Ipkabs⋅

le69.967=:=

Lb_FL 0.61 Ntf2

⋅ AL⋅ 103−

⋅ 23.256=:=

Lb_NL Ntf2

AL⋅ 103−

⋅ 38.125=:=

117

mΩ/cm

Wire Area cm^2

Peak Inductor Current

Min Inductor Current

Max DC Magnetizing force

Min DC Magnetizing force

From the following curve for 60u Material:

Maximum Flux Density Gauss

Minimum Flux Density Gauss

Rpu 0.1319:=

Wa 0.0152:=

Ff NtfWa

Aw⋅ 0.244=:=

Ipk_max Ipk 0.5 ∆I Lb_LL⋅+ 14.142=:=

Ipk_min Ipk 0.5 ∆I Lb_LL⋅− 8.485=:=

Hor_max0.4 π⋅ Ntf⋅ Ipk_max⋅

le69.967=:=

Hor_min0.4 π⋅ Ntf⋅ Ipk_min⋅

le41.98=:=

Bmax 3800:=

Bmin 2500:=

∆B Bmax Bmin−:=

1

2∆B⋅ 650=

118

Core losses mW/cm^3

Core losses W

Coil resistivity mΩ

DC Copper Losses W

Wire Gage factor for AWG#16

AC Resistance at Switching frequency

AC Copper Losses

Total Inductor losses per phase

PFC Bus Capacitor Selection:

Capacitance Value

Minimum Hold Up time for 50Hz

Maximum Allowable PFC Voltage Droop

PFC DC Current

Minimum Required Bus Capacitance

Pc_pu1

2

∆B

103

2fs

1000

1.46

⋅ 795.54=:=

Pc Pc_puVe

1000⋅ 3.301=:=

Rdc MLT Ntf⋅ Rpu⋅ 17.246=:=

PcuRdc

1000IL

2⋅ 1.104=:=

Kf 14.25:=

RacRdc

100Kf⋅

fs

106

⋅ 1.028=:=

Pac Rac∆I Lb_LL

2 3⋅

2

⋅ 2.742=:=

Ptot Pc Pcu+ Pac+ 7.147=:=

Tholdup1

8

1

50⋅ 2.5 10

3−×=:=

Vpfc_droop 20:=

IoPout1

Vboost9.062=:=

Cpfc2 Pout1⋅ Tholdup⋅

Vboost2

Vboost Vpfc_droop−( )2

:=

Cpfc 1.162 103−

×=

119

Capacitors RMS Current:

Total rms ripple current of PFC Caps

2 x Line Frequency rms ripple current

High Frequency Ripple current of PFC Caps

Part 2 - Control Component Selction

Device Parameters Definition

Refrence and Internal Bias Voltage

Current Amplifier Ox Transconductance

Voltage Amplifier O Transconductance

PWM ramp amplitude

Timing regulation Voltage at RDM

Timing regulation Voltage at RT

Timing regulation Voltage at DMAX

1 - Component Selection for RT and DMX:

Calculate RRT for target switching frequency: target

Standard Value chosen

ichf_t Io16 Vboost⋅

6 π⋅ 2⋅ Vinmin⋅ η12

1−⋅ 12.528=:=

ichf_lfIo

26.408=:=

ichf_hf Io16 Vboost⋅

6 π⋅ 2⋅ Vinmin⋅ η12

1.5−⋅ 10.765=:=

VREF 6:=

gmi 100 106−

⋅:=

gmv 70 106−

⋅:=

Vramp 4:=

VRDM 3:=

VRT 3:=

VDMAX 3:=

fsw 175:= KHz

RRT7500 10

3⋅

fsw4.286 10

4×=:=

RRT 43.2 103

⋅:=

120

Actual nominal switching frequency (KHz) will be:

Power dissipation is negligible. Select a suitable size and power rating for the resistor.

Calculate RRDM for desired maximum duty cycle:

Target

Standard Value chosen

Actual

2 - Component Selection for CDR and RDM frequency Dithering:

KHz Frequency Dithering Range

Standard Value chosen

KHz Frequency Dithering Range

pF

pF

3 - Component Selection - Vsense snd Vinac Resistor Configuration:

Choose the voltage feedback resistor divider network: VREFfb = 3

desired target

fsw750010

3⋅

RRT

:= fsw 173.611= KHz

Dmax 0.95:=

RDMX RRT 2Dmax 1−( )⋅:= RDMX 3.888 104

×=

RDMX 39.2 103

⋅:=

Dmax1

2

RDMX

RRT

1+

:= Dmax 0.954=

fdm 0.07 fsw⋅ 12.153=:=

RRDM937.510

3⋅

fdm7.714 10

4×=:=

RRDM 76.8 103

⋅:=

fdr 1:=

CRDM 66.7

RRDM

1000

fdr

⋅ 5.123 103

×=:=

CRDM 5100:=

VREFfb 3:= VOUT_AVG 400:=

121

Voltage divider resistor B

RA = R1 + R2 + R3 Voltage divier resistors A

Total power dissipation

Select Components for Vinac Divider:

Vinac sense resistors must be the same divider-ratio as that of the Vout sense resistors. Perferably, they would be the same values and ratings.

4 - Size current-sense resistor for a 3.0V dynamic range:

Choose current sense transformer turns ratio NCT:

(Use 120% of Ipk to allow for a little extra current/power to recover from large load-step or line-step changes, or line drop-outs.)

Choose a standard value:

(Power will actually be somewhat lower than this because of less than 100% duty-cycle of Ipk.)

Choose a 1/4-watt resistor.

RB 24.9 103

⋅:=

RA RB

VOUT_AVG

VREFfb

1−

⋅ 3.295 106

×=:=

R1 1.1 106

⋅:=

RA 3 R1⋅:=

PRdiv

VOUT_AVG2

RA RB+:= PRdiv 0.048=

VOUT

RA

RB

1+

VREFfb⋅:= VOUT 400.59=

NCT 100:=

Rs3 NCT⋅

Ipk 1.20⋅:=

Rs 22.097=

Rs 22.1:=

P_rsIpk

2

2Rs

NCT2

⋅:=

P_rs 0.141=

122

5 - Select components for Pulse by Pulse Current limiting:

a. Calculate the peak limit voltage: Peak current limit is a cycle-by-cycle limit which determines the absolute peak value of MOSFET drain current allowable (per phase). If this limit is exceeded, the current PWM cycle is shut down. The peak value of the inductor ripple current must be included, as well as any multipling factor used in the average current limit calculation.

b. Calculate Ipeak Resistor Divider

Limit current out of Vref to less than 0.5mA to minimize loading, since these resistors will be off of the Vref.

Select standard bottom divider-resistor value

Select standard top divider-resistor value

Resistor power dissipation is negligible; choose any suitable size and rating.

Actual peak limit voltage:

Actual peak MOSFET current per phase:

VpklmtIpk 0.5 ∆I Lb_LL⋅+( ) 1.20⋅

NCT

Rs⋅:= Vpklmt 3.75=

VREF

0.5 103−

1.2 104

×= VREF 0.5 103−

⋅( )⋅ 3 103−

×=

Vpklmt

VREF

1.2 104

×( )⋅ 7.501 103

×=

Rpklmt2 7.5 103

⋅:=

Rpklmt1VpklmtRpklmt2⋅

VREF

:=

Rpklmt1 4.688 103

×=

Rpklmt1 4.53 103

⋅:=

Vpklmt VREFRpklmt2

Rpklmt1 Rpklmt2+⋅:= Vpklmt 3.741=

IpeakVpklmt

RsNCT⋅:= Ipeak 16.926=

123

Make sure PKLMT voltage is set properly to: a) co-ordinate with low-line power limit, b) account for PKLMT detection propagation delay, and c) allow for additional saturation swing at extra power levels.

6 - Current Synthesizer:

The Resistor-Divider Atenuation at the Vsense and Vinac

Boost Inductor uH

KΩ Selected Standard Resistor

7 - Multiplier Set up:

Output of Voltage-error Amplifier

Min Voltage Feed Forward Factor

Lowest Max Power Limit

Accounting for 2-V bridge drop

Choose standard value:

kR

RB

RA RB+:=

LB Lb 106

⋅ 86=:=

RSYN

10 NCT⋅ LB⋅ kR⋅

Rs29.143=:=

RSYN 30.1:=

VAOmax 5:=

kVFF 0.398:=

VVINAC 0.76:=

Imo_max

17 106−

⋅ VVINAC⋅ VAOmax 1−( )⋅

kVFF

:= Imo_max 1.298 104−

×=

VINAC 73:=

Iin_pk 2Pout1 1.1⋅

η1⋅

1

VINAC

⋅:= Iin_pk 78.421=

RIMOIpkabs Rs⋅

NCT Imo_max⋅:= RIMO 2.407 10

4×=

RIMO 24.9 103

⋅:=

124

8 - Current Loop Compensation Design:

For I-loop compensation calculations, set Lboost to mid-way between max and min swing values (50uH to ~100uH).

select target cut off frequency

Current-Loop Plant Transfer Function:

Initial I-loop compensation target values:

i 1 2000..:=

fi

100

i 200−( )

500:=

wi

fi

:=

si

2 π⋅ j⋅ wi

⋅:=

Vout 400:=

Lboost 86 106−

⋅:=

fcxo 16 103

⋅:=

Gipi

VoutRs

NCT

si

Lboost⋅ Vramp⋅:=

dBGipi

20 log Gipi( )⋅:= φGip

iatan2 Re Gip

i( ) Im Gipi( ), ( )

180

π⋅:=

Rz1

gmi

VoutRs

NCT

2 π⋅ fcxo⋅ Lboost⋅ Vramp⋅

3.912 103

×=:=

Cz1

2 π⋅fcxo

3

⋅ Rz⋅

7.628 109−

×=:=

Cp1

2 π⋅ 2 fcxo⋅( )⋅ Rz⋅1.271 10

9−×=:=

125

(Note: values for Rzi, Czi, Cpi for the following GCEA equation are entered as global

variables defined on the next page, below.)

Compensator Transfer Function

Rz 3.912 103

×=

Cz 7.628 109−

×=

Cp 1.271 109−

×=

Rzi 4 103

⋅≡

Czi 75001012−

⋅≡

Cpi 12001012−

⋅≡

GEAi

gmi 1 si

Rzi⋅ Czi⋅+( )⋅

si

Czi Cpi+( )⋅ 1 si

Rzi⋅Cpi Czi⋅

Cpi Czi+

⋅+

:=

dBGEAi

20 log GEAi

⋅:=φG EA

iatan2 Re GEA

i

Im GEAi

,

180

π⋅:=

10 100 1 103

× 1 104

× 1 105

× 1 106

× 1 107

×

40−

26.667−

13.333−

0

13.333

26.667

40

180−

150−

120−

90−

60−

30−

0

Current Loop Plant & Compensator Transfer Function

dBGipi

dBGEAi

φGip i

φG EAi

f i

126

Current Loop Gain (open-loop transfer function):

Vary parameters to see how crossover is affected:

Phase margin (PM) should be ~ 45deg at the cross-over frequency fcxo as

parameters vary.

GOLi

VoutRs

NCT

si

Lboost⋅ Vramp⋅

gmi 1 si

Rzi⋅ Czi⋅+( )⋅

si

Czi Cpi+( )⋅ 1 si

Rzi⋅Cpi Czi⋅

Cpi Czi+

⋅+

:=

dBGOLi

20 log GOLi

⋅:=

φG OLi

atan2 Re GOLi

Im GOLi

,

180

π⋅:=

PM_GOLi

φG OLi

180+:=

100 1 103

× 1 104

× 1 105

× 1 106

×

100−

80−

60−

40−

20−

0

20

40

60

80

100

180−

126−

72−

18−

36

90

Current Open Loop Bode Plot

dBGOLi

φG OLi

PM_G OLi

f i

127

9 - Voltage Loop Compensation Design:

a. Evaluate Modulator and Power Stage Gain

PFC Voltage-Loop Boost Plant Transfer Function:

This approximation is valid for frequencies well below the LC resonance frequency. The low voltage-loop cross-over freq helps ensure this condition is true.

PinPout1

η13.68 10

3×=:=

∆Vea 3.3:=

Cpfc 1200106−

⋅:=

Gvpi

Pin1

∆Vea Cpfc⋅ Vout⋅ si

⋅:=

dBGvpi

20 log Gvpi

⋅:= φG vpi

atan2 Re Gvpi

Im Gvpi

,

180

π⋅:=

0.1 1 10 100 1 103

×

100−

80−

60−

40−

20−

0

20

40

60

80

100

135−

117−

99−

81−

63−

45−

Voltage Loop Plant Transfer Function

dBGvpi

φG vpi

f i

128

b. Evaluate Compensation Scheme

H1 is "gain" of the VSENSE resistor-divider network.

From above, Bulk Capacitor zero-peak ripple voltage:

for

Voltage error amplifier gain is adjusted with compensation components to attenuate the twice-line frequency ripple (vClf) on the output capacitor to obtain a desired reduction of 3rd harmonic THD.

Cpv sets the reduction level, Rzv sets the phase margin to 45degrees at fvxo, and Czv sets the beginning of phase boost.

Vary standard values and parameters to check fvxo and Phase Margin (PM) under different conditions. Typically: 0.00 < kPin < 1.0, 0.8 < kCout < 1.2, 0.8 < kgmv < 1.2

VREFfb 3:= Vout 400=

H1

VREFfb

Vout:= H1 7.5 10

3−×=

1

H1

133.333=

vClfPin1

2 π⋅ 100⋅ Cpfc Vout⋅:= vClf 12.202= Cpfc 1.2 10

3−×=

%k3rd 1.5:= f2LF 100:= gmv 70 106−

⋅:=

Zov f2LF( )%k3rd 66⋅ 10

3−× Vout⋅ 2⋅ π f2LF Cpfc⋅

gmv H1⋅ Pin1⋅:=

Cpv1

2π f2LF Zov f2LF( )⋅:= Cpv 1.03 10

7−×=

fvxo

gmv H1⋅ Pin1⋅

∆Vea Vout⋅ 2π( )2

⋅ Cpv⋅ Cpfc⋅

:= fvxo 17.321=

Rzv1

2π fvxo Cpv⋅:=

Rzv 8.923 104

×=

Czv10

2π fvxo Rzv⋅:= Czv 1.03 10

6−×=

129

Calculated values:

Select standard values:

Cpv 1.03 107−

×= Rzv 8.923 104

×= Czv 1.03 106−

×=

Cpv 0.05 106−

⋅:= Rzv 91 103

⋅:= Czv 0.5 106−

⋅:=

GVLi

Pin1

∆Vea Cpfc⋅ Vout⋅ si

⋅H1⋅

01.0 gmv⋅( ) si

Rzv⋅ Czv⋅ 1+( )⋅

si

Czv Cpv+( )⋅

si

Rzv⋅ Czv⋅ Cpv⋅

Czv Cpv+1+

⋅:=

dBGVLi

20 log GVLi

⋅:=φG VL

iatan2 Re GVL

i

Im GVLi

,

180

π⋅:=

PM_GVLi

φG VLi

180+:=

0.1 1 10 100 1 103

×

100−

80−

60−

40−

20−

0

20

40

60

80

100

180−

126−

72−

18−

36

90

Voltage Open Loop Bode Plot

dBGVLi

φG VLi

PM_G VLi

f i

130

Part 3 - Load Step Change Transient Response

Load Step Change Transient Response from 5% to 95% of Full Load

Top: Output Voltage Bottom: Input Current

Load Step Change Transient Response from 95% to 5% of Full Load

Top: Output Voltage Bottom: Input Current