Top Banner
A CASCADE BOOST CONVERTER DESIGN, DEMONSTRATION, AND SCALING FOR FUTURE HIGH VOLTAGE POWER CONDITIONING SYSTEMS A Thesis presented to the faculty of the Graduate School University of Missouri—Columbia In Partial Fulfillment Of the Requirements for the Degree Master of Science by SCOTT CASTAGNO Dr. Randy Curry, Thesis Supervisor MAY 2006
149
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Research

A CASCADE BOOST CONVERTER DESIGN, DEMONSTRATION, AND SCALING FOR FUTURE HIGH VOLTAGE POWER CONDITIONING SYSTEMS

A Thesis presented to the faculty of the Graduate School

University of Missouri—Columbia

In Partial Fulfillment

Of the Requirements for the Degree

Master of Science

by

SCOTT CASTAGNO

Dr. Randy Curry, Thesis Supervisor

MAY 2006

Page 2: Research

© Copyright by Scott Castagno 2006

All Rights Reserved

Page 3: Research
Page 4: Research

ii

ACKNOWLEDGEMENTS

I would like to thank Dr. Randy Curry for his assistance, guidance, and support

with the research and writing of this thesis. All of his help and dedication is greatly

appreciated. Additionally, I would like to recognize Dr. Kenneth McDonald for his

willingness to provide technical support for the work. Also, student colleagues Josh

Leckbee, Peter Norgard, Laura Heffernan, Keith Lechien, and Matt Aubuchon provided

valuable collaboration and assistance throughout my graduate studies.

Page 5: Research

iii

TABLE OF CONTENTS

ACKNOWLEDGEMENTS................................................................................................ ii

TABLE OF CONTENTS................................................................................................... iii

LIST OF TABLES............................................................................................................ vii

LIST FIGURES .................................................................................................................. x

CHAPTER 1.0 INTRODUCTION ..................................................................................... 1

REFERENCES FOR CHAPTER 1.0 ................................................................................. 9

CHAPTER 2.0 THEORETICAL CASCADE BOOST DESIGN .................................... 11

2.1 SWITCH-MODE CONVERTERS................................................................. 11

2.2 THE BOOST CONVERTER.......................................................................... 13

2.3 DISCONTINUOUS MODE ........................................................................... 17

2.4 CASCADE BOOST CONVERTER DESIGN ............................................... 21

2.4.1 VOLTAGE GAIN AND INDUCTOR DERIVATION................... 22

2.4.2 FILTER CAPACITORS .................................................................. 27

2.5 ASYMMETRICAL SWITCHING ................................................................. 29

Page 6: Research

iv

2.6 INITIAL DESIGN VALUES AND SIMULATION RESULTS OF AN

IDEALIZED MODEL .......................................................................................... 31

REFERENCES FOR CHAPTER 2.0 ............................................................................... 37

CHAPTER 3.0 CASCADE BOOST CONVERTER PROTOTYPE DESIGN................ 38

3.1 IGBT SWITCHING DESIGN THEORY....................................................... 39

3.1.1 IGBT CAPACITANCE ................................................................... 41

3.1.2 THE TURN-OFF TRANSIENT OF IGBTS ................................... 44

3.1.3 IGBT SWITCHING THEORY CONCLUSION............................. 56

3.2 FIRST STAGE DETAIL ................................................................................ 57

3.3 SECOND STAGE DETAIL ........................................................................... 62

3.4 SECOND STAGE SWITCH DESIGN OPTIMIZATION............................. 68

REFERENCE FOR CHAPTER 3.0.................................................................................. 73

CHAPTER 4.0 DIAGNOSTICS OF THE CASCADE BOOST CONVERTER

PROTOTYPE ................................................................................................................... 75

4.1 VOLTAGE DIAGNOSTICS.......................................................................... 75

4.2 CURRENT DIAGNOSTICS .......................................................................... 77

4.3 OSCILLOSCOPE ........................................................................................... 78

Page 7: Research

v

CHAPTER 5.0 CASCADE BOOST CONVERTER PROTOTYPE RESULTS ............. 79

5.1 CASCADE BOOST CONVERTER OUTPUT AND EFFICIENCY ............ 80

5.2 COMPONENT LOSS ANALYSIS ................................................................ 88

5.2.1 MEASUREMENT OF THE FIRST STAGE IGBT LOSS ............. 88

5.2.2 MEASUREMENT OF THE SECOND STAGE IGBT LOSS ........ 92

5.2.3 MEASUREMENT OF DIODE LOSSES ........................................ 94

5.2.4 TOTAL COMPONENT LOSS CALCULATION .......................... 98

5.3 OPTIMIZATION OF THE PROTOTYPE CONVERTER............................ 99

REFERENCES FOR CHAPTER 5.0 ............................................................................. 101

CHAPTER 6.0 CASCADE BOOST CONVERTER SCALING TO HIGHER POWER

LEVELS.......................................................................................................................... 102

6.1 NEAR-TERM TECHNOLOGY BASED SCALING .................................. 103

6.2 FIVE & TEN-YEAR FUTURE TECHNOLOGY SCALING ..................... 112

REFERENCE FOR CHAPTER 6.0................................................................................ 123

CHAPTER 7.0 CONCLUSION...................................................................................... 125

APPENDIX A................................................................................................................. 128

Page 8: Research

vi

APPENDIX B ................................................................................................................. 131

APPENDIX C ................................................................................................................. 132

APPENDIX D................................................................................................................. 134

Page 9: Research

vii

LIST OF TABLES

Table

2.1 Comparison of voltage gain and inductor values over a range of duty ratios and voltage gain ratios. DCM inductor held at 3.3 mH; CCM Lmin adjusted to keep boost converter in CCM. .................................................................................................... 19

3.1 Cascade boost converter prototype design goals. ...................................................... 39 3.2 Caddell-Burns inductor specifications. Incremental current is the minimum current

at which the inductance will be decreased by 5% from the initial (zero-DC) value. 60 3.3 Volume and mass of first stage components using COTS components. ................... 61 3.4 Volume and mass of the second stage components. .................................................. 67 3.5 Total cascade boost converter prototype volume and mass....................................... 68 3.6 Turn-off loss comparison of the two IGBT designs. ................................................. 71 3.7 Cascade boost converter parameter comparison of 3.3 kV and 1200 V IGBTs. ....... 72 4.1 Voltage probes for circuit diagnostics including attenuation factor and bandwidth. 76 4.2 Measurement and diagnostic equipment implemented.............................................. 77 4.3 Current monitors for circuit diagnostics including output voltage per amp, peak

current, max continuous current, usable rise time. ................................................... 78 5.1 Cascade boost converter prototype design goals. ...................................................... 79 5.2 Average value data of the system operated at 5 kV output........................................ 84 5.3 Efficiency and loss for the first and second stages operated at 5 kV output. ............ 84

Page 10: Research

viii

5.4 First stage and second stage voltage data of the cascade boost converter prototype as input voltage is varied from 0 V to 100 V. ............................................................... 84

5.5 Cascade boost converter operating conditions for the component loss analysis ....... 88 5.6 Summary of the first stage IGBT data. ...................................................................... 92 5.7 Summary of second stage IGBT data. ....................................................................... 94 5.8 Breakdown of average component power loss. ......................................................... 99 6.1 Cascade boost converter specifications for a 60 kV, 300 kW module. ................... 102 6.2 Ideal parameters calculated for a 60 kV, 300 kW cascade boost converter. ........... 104 6.3 Switching parameters adjusted, corresponding to 90% efficiency. ......................... 104 6.4 Number of IGBT modules in series and parallel for assembly of S1 and S2............ 106 6.5 Expected voltage and current per IGBT, and total average turn-off loss. ............... 106 6.6 First stage components volume and mass for the near-term 60 kV, 300 kW converter.

................................................................................................................................. 110 6.7 Second stage components volume and mass for the near-term 60 kV, 300 kW

converter. ................................................................................................................ 110 6.8 Estimated volume and mass of components of the near-term 60 kV, 300 kW

converter. ................................................................................................................ 111 6.9 Total near-term projected component volume and mass and complete system volume

and mass of a 300 kW converter. The 20% scaling factor estimates additional volume and mass for a completely packaged and operational converter................ 111

6.10 Properties of SiC that will enhance power semiconductor devices. ...................... 113 6.11 Summary of the key parameter estimates for SiC devices in 5 and 10 years. ....... 116 6.12 Number of devices to be placed in series and parallel for the 5 and 10-year time

frames for the 300 kW converter’s first and second stage switches. ...................... 116 6.13 First stage component projected volume and mass for the 5 & 10-year trend in a 60

kV, 300 kW converter module................................................................................ 119

Page 11: Research

ix

6.14 Second stage component projected volume and mass for the 5 & 10-year trend in a 60 kV, 300 kW converter module........................................................................... 119

6.15 Estimated total volume and mass of each component category in a 60 kV, 300 kW

converter module. ................................................................................................... 119 6.16 5-year total projected component and complete system volume and mass for a 60

kV, 300 kW converter module. The 20% scaling factor estimates additional volume and mass for a completely packaged and operational converter............................. 119

6.17 10-year total projected component and total system volume and mass for a 60 kV,

300 kW converter module. The 20% scaling factor estimates additional volume and mass for a completely packaged and operational converter. .................................. 120

Page 12: Research

x

LIST OF FIGURES

Figure

1.1 Schematic of a simplified cascade boost converter. .................................................... 4 1.2 CPI gyrotron cross section. .......................................................................................... 6 1.3 CPI depressed collector configuration. ........................................................................ 7 2.1 General diagram of a generic DC-DC switch-mode converter.................................. 12 2.2 Single Stage Boost Converter .................................................................................... 14 2.3 Discontinuous Inductor Current. D is the switch duty ratio and ∆1 is the diode duty

ratio or the fall time of the inductor current.............................................................. 15 2.4 Continuous Inductor Current. During the fall time of the inductor current, the current

remains greater than zero. ......................................................................................... 15 2.5 Comparison of DCM to CCM (R=25 kΩ, f=10 kHz, L=3.3 mH)............................. 19 2.6 The CCM diode and inductor currents. High dif/dt occurring at diode turn-off....... 21 2.7 The DCM diode and inductor currents ...................................................................... 21 2.8 Simplified Cascade boost converter schematic.......................................................... 22 2.9 Example of the current waveforms of L1 (red), L2 (blue), and C1 (black). D is the

switch duty ratio, and ∆1 and ∆2 are the duty ratios of the diodes and fall time of the inductor current......................................................................................................... 23

2.10 Idealized Pspice circuit model using derived component values. Resistors R1, R2,

R3 were inserted into the model to avoid convergence problems. ........................... 33 2.11 Simulated output voltage of the second stage/converter. ∆V = 47.7 V, which

correlates to a voltage ripple of less than ± 0.5%..................................................... 34

Page 13: Research

xi

2.12 Simulated voltage across C1 and current through L1.............................................. 35 2.13 Simulated current through L2 with the second stage output voltage across C2. ..... 35 3.1 Bench top cascade boost converter prototype............................................................ 39 3.2 The IGBT equivalent circuit including, internal MOSFET and BJT, switch

capacitance, and drift region resistance, Rdrift. .......................................................... 42 3.3 The IGBT turn-off waveforms in a clamped inductive load. td(off): dealy time; tvr:

voltage rise-time; tf1: 1st current fall-time; tf2: 2nd current fall-time.......................... 42 3.4 Cross-section of the NPT IGBT................................................................................. 47 3.5 PNP transistor layers in the IGBT: Epeak occurs at the junction J2, and the depletion

region width, Wd, extends primarily into the n- drift region due to high doping concentration of the p+ body region. The x axis is the distance along the pnp layers, where the zero occurs at junction J2. ........................................................................ 47

3.6 Schematic of first stage components......................................................................... 57 3.7 Thermaflo E3045 heat sink profile. ........................................................................... 59 3.8 Schematic of second stage components..................................................................... 62 3.9 The IGBT stack with six IGBTs in series, operating as the second stage switch...... 64 3.10 The output Load with 10 series Ohmite, wire wound resistors for dissipation of the

1 kW output of the prototype. ................................................................................... 67 3.11 Voltage and current waveforms of two 3.3 kV IGBTs in series as the second stage

switch of the cascade boost converter with and output of 5 kV and 1 kW............... 70 3.12 Power loss waveform of the two 3.3 kV IGBTs in series. The integrated energy

loss is 126.5 mJ per pulse, which results in an average power loss of 1139 W........ 70 5.1 First stage DC output voltage; the average output voltage is displayed at 503 V, for a

voltage gain slightly of 5 from the 100 V input........................................................ 81 5.2 Cascade boost DC converter output voltage; the average output voltage is displayed

at 5010 V, for a voltage gain of approximately 10 from the 503 V first stage output.................................................................................................................................... 82

Page 14: Research

xii

5.3 First stage and second stage output voltages as input voltage is varied from 0 to 100 V, with constant switching duty ratios of .614 and .513 for the first and second stages respectively. ................................................................................................... 85

5.4 Efficiency of the individual stages and the overall converter.................................... 85 5.5 First stage experimental and simulated input currents when the converter is operated

with a 100 V input and 5 kVoutput........................................................................... 87 5.6 Second stage experimental and simulated input currents when the converter is

operated with a 100 V input and 5 kV output. .......................................................... 87 5.7 Turn-off voltage and collector current waveforms of the first stage IGBTs in parallel.

................................................................................................................................... 91 5.8 First stage turn-off power loss. The integrated turn-off loss = 4.83 mJ at 20 kHz for

98.57 W average loss. ............................................................................................... 91 5.9 Turn-off voltage and current waveforms of the second stage IGBT stack. ............... 93 5.10 Second stage IGBT stack turn-off power loss. The integrated turn-off energy loss =

11.6 mJ at 9 kHz for 104.7 W average loss. ............................................................. 93 5.11 First stage diode voltage and current waveforms. ................................................... 95 5.12 First stage diode power loss waveform. Integrated energy loss is 0.82 mJ per pulse

and average power loss is 1.67 W............................................................................. 96 5.13 Second stage diode voltage and current waveforms. ............................................... 97 5.14 Second stage diode power loss waveform. Integrated energy loss is 0.457 mJ per

pulse and average power loss is 4.13 W. .................................................................. 97 6.1 300 kW converter volume scaling from estimated technology trends..................... 121 6.2 300 kW converter mass scaling from estimated technology trends......................... 121 6.3 1.2 and 2.4 MW converter volume scaling using parallel 300 kW converter modules.

................................................................................................................................. 122 6.4 1.2 and 2.4 MW converter mass scaling using parallel 300 kW converter modules.

................................................................................................................................. 122 A.1 IGBT trigger module........................................................................................................129

Page 15: Research

xiii

A.2 Power oscillator....................................................................................................... 129 A.3 IGBT gate driver circuit.......................................................................................... 130 C.1 Experimental on-state forward conduction voltage and trend line of the International Rectifier IRGP30B120KD-E 1200 V, 60 A rated, Non-Punch Through (NPT),IGBT; Vg = 12.5 V. ........................................................................................................................ 133

Page 16: Research

1

CHAPTER 1 .0 INTRODUCTION

The demand for compact electrical systems has motivated size and weight

reduction of power conditioning architecture sub-systems. For military applications, the

government has funded the development of compact high voltage electrical systems for

combat vehicles that require large scale power conditioning. These systems include

electromagnetic launchers, electromagnetic armor, and electronic weapon systems [1-2].

These technologies have the potential for revolutionizing military engagement. However,

they face engineering challenges to meet the size and weight requirements for integration

on mobile platforms, in addition to technical risk mitigation and adhering to the stringent

reliability requirements for operation in military environments. The power conditioning

becomes a significant factor of the total size and weight of these systems, where new

technologies and innovation must develop highly compact and lightweight conditioning

architectures in order for such emerging military technologies to move forward.

The development of power conditioning technology has led to advancements in

the area of switch-mode power converters [3-4]. Switch-mode power conversion

topologies are highly common among electronic devices and power conditioning designs.

A switch-mode converter circuit topology familiar to the subject of power electronics is

the step-up, or boost converter [5-7]. The boost converter is a DC to DC converter that

simply steps-up, or boosts, the input voltage to a higher output voltage. Utilizing the

same circuit operation, AC to DC boost converters are used as power factor correction

(PFC) circuits for the front end of power conditioning systems requiring a unity power

Page 17: Research

2

factor interface with AC power sources. Generally, commercial boost converter circuits

are implemented in systems requiring power conditioning levels ranging from less than

10 W to multi-kilowatt, and have output DC voltages in the hundreds of volts range or

less, with step-up ratios on the order of 10. Converter designs in this regime have the

ability to exceed 90% efficiency [8-10]. The general operation principles of switch-mode

converters and typical boost converter design equations are discussed in this thesis, as a

background to the design of the cascade boost converter.

In converter designs for high voltage applications, high voltage switching is a

significant design challenge. Solid-state components rated higher than 1 kV can limit

switching frequency and efficiency due to typically higher switching loss, which impacts

component size, thermal management, and total power density. Fortunately, for solid-

state switch technology the capability of high voltage solid-state devices is advancing.

The pulsed power industry also has taken advantage of this technology as solid-state

switching provides advantages such as increased reliability, lifetime, control, and

provides the ability to create more compact system designs. For example, high voltage

thyratrons and spark gaps have been replaced by solid-state switches, and high voltage

power modulator designs have taken advantage of the solid-state technology [11-14].

Similarly, switch-mode converters and high voltage power conditioning systems benefit

with higher switching frequencies, greater power densities, and higher efficiencies from

improved, emerging solid-state switch technology. For instance, wide band-gap

semiconductor materials like silicon carbide (SiC) offer advantages, over existing silicon

(Si) devices, such as higher electric field breakdown strength, faster switching, and lower

Page 18: Research

3

switching loss. Already, 1200 V SiC diodes exhibit lower loss and can allow converters

to operate at higher frequencies. High voltage diodes with a voltage rating of 10 kV or

higher are anticipated to emerge as the first generation of high voltage SiC devices with

5-10 kV rated transistors subsequently available. A semiconductor paradigm shift of

mature wide band-gap devices for high voltage switching may occur within a 10-year

time frame. In addition to semiconductors, developing high voltage capacitor dielectrics

are enabling capacitive energy storage in smaller volumes, and nanocrystalline magnetic

core materials have already shown improvements in transformer and inductor core

reduction.

The circuit topology investigated in this research project is a high voltage

transformerless DC-DC switching converter, which has a significant potential for

reduction in size and weight through advances in the aforementioned SiC, capacitor, and

nanocrystalline technologies. The anticipated SiC technology will allow the converter to

operate at higher switching frequencies, thus shrinking passive components. Also,

expanded voltage capability will reduce the number of series devices in stacked assembly

designs. This converter is composed of two series arranged boost converters and

develops a high input to output DC-DC voltage step-up gain from the two stages. The

topology will suitably be referred to as the cascade boost converter, and a simplified

schematic is shown below in Figure 1.1.

Page 19: Research

4

R L

D 2 L 2 L 1 D 1

C 2 S 2 C 1 S 1 V s

Figure 1.1 Schematic of a simplified cascade boost converter.

A particular application for which the cascade boost converter is designed is the

power conditioning for high power microwave tubes. The Department of Defense

(DOD) has developed microwave architectures for integration onto airborne, sea, and

land vehicles. The Air Force has accelerated research and development of Directed

Energy systems using high-power microwave and laser technology, and this effort is

expected to continue well into the twenty-first century [15]. These Directed Energy

weapon systems can provide non-lethal tactical solutions for the military such as remote

disruption or destruction of enemy electronic systems and even physically stunning

enemy combatants with applied radiation on the subjects. Directed Energy weaponry

provides the United States with means for tactical advantage over the country’s

adversaries where casualty mitigation is an objective. The cascade boost topology is a

potential enabling technology for systems for mobile platforms by fulfilling power

conditioning size requirements. However, this converter is certainly not limited to high-

power microwave systems. The output can be tailored to meet the needs of a wide

variety of applications.

A typical high power microwave source is the gyrotron microwave tube.

Gyrotrons, originally developed in the 1960’s, were utilized for electron cyclotron

resonance heating of plasmas in fusion experiments [15]. Since the inception of the

Page 20: Research

5

Gyrotron, consortiums in Russia, France, Japan, and United States have developed

megawatt continuous power class tubes [15]. In a Gyrotron, a cathode at modest 40-85

kV voltage levels injects an electron beam into a 2-3 cavity structure [16]. The

continuous electron beam, after passing through the first cavity, produces a periodically

bunched beam [16]. An external magnetic field is also applied, imparting a helical

motion to the electron beam [16]. This helical motion allows the coupling of the electron

beam to a fast wave structure for extraction of the microwave power [16]. Gyrotron

tubes generate high power and high frequency microwave radiation that is applicable for

Directed Energy systems.

The predominant manufacturer of Gyrotrons in the United States, Communication

and Power Industries (CPI), has manufactured Gyrotrons up to 1 MW power levels at

efficiencies of 20-45% [17]. The typical efficiency of a single extraction cavity Gyrotron

is 20%, while multiple extraction cavities exhibit efficiencies of 40-45%. Gyrotrons with

an extraction efficiency of 40-45% and megawatt power levels are under development

[17].

The Gyrotron consists of a cathode, a ceramic metal housing, superconducting

magnets, a vacuum pump (VacIon) and a collector. A cross section of a CPI gyrotron is

shown in Figure 1.2. A 60 kV, 40A power supply is utilized to power the megawatt

output Gyrotron. In general, mega-watt class gyrotron tubes require the electron gun

cathode power supply to deliver 20 A or greater [18]. The body of the tube is biased at

+20 kV in the depressed collector mode of operation, where the collector is grounded

(depressed in voltage relative to the body) as shown in Figure 1.3. The 20 kV body

Page 21: Research

6

supply is a 50mA, highly regulated power supply. The 20 kV power supply is typically

regulated to +/- 0.1%. The efficiency of the Gyrotron is somewhat determined by the

regulation of the power supply according to CPI. Thus, the 60 kV supply must also have

a minimum regulation of +/- 0.5%.

Figure 1.2 CPI gyrotron cross section.

Page 22: Research

7

Figure 1.3 CPI depressed collector configuration.

The cascade boost converter has been selected to operate at 60 kV and 300 kW at

full power, based on the high current gyrotron power supply requirements, with the

ability to be modularized for higher power conditioning output by paralleling modules.

This modular approach extends the flexibility for many applications and provides

redundancy for increased reliability. Power regulation (+/- 0.5%) is achieved with

appropriate output filtering and a precise boost converter control system. The input for

the design of the converter has been selected at 1 kV, which requires a 60 times step-up

ratio for a 60 kV converter output.

Boost converter theory, an idealized converter design, an experimental cascade

boost converter prototype for design validation, and theoretical system scaling to mulit-

megawatt output levels is presented in the following chapters. The prototype converter

has demonstrated an output of 5 kV at 1 kW, stepped up from a 100 V input, for a total

step-up ratio of 50. The load utilized in the tests is a resistive load, and the converter was

designed for analysis and circuit demonstration in the laboratory environment, where the

(2) + 60 kV 40 A

_ Voltage Divider

for Sensing

(1) - 20 kV 50 mA

+

Collector

Body

1– Low current, highly regulated supply (+/- 0.1% regulation) 2– High current, moderately regulated supply (+/- 0.5% regulation)

Gyrotron

Page 23: Research

8

converter was operated at a constant input and output voltage. The scope of the prototype

converter analysis is centered on the main boost circuit components. This analysis

includes individual component power loss and performance in relation to the system

performance, namely the switches, diodes, inductors, and capacitors of the converter.

This research project also includes the review of high voltage solid-state switching theory

and switching designs of the cascade boost converter. Finally, the cascade boost

converter prototype is used to estimate the volume and mass of a 60 kV, 300 kW

converter designed for near-term, 5-year, and 10-year projected technology trends. The

scaling trend of the cascade boost converter demonstrates the dramatic reduction possible

for such a system, and projects the potential for the compact power conditioning systems

based on the topology of the cascade boost converter.

Page 24: Research

9

REFERENCES FOR CHAPTER 1.0 [1] J.A. Gaudet, “Research issues in developing compact pulsed power for high peak

power applications on mobile platforms,” Proceeding of the IEEE, vol. 92, No. 7, pp. 1180-1196, July 2004.

[2] I.R. McNab, “Developments in battlefield power technology,” in Proc. of Pulsed Power Conf., 1999, pp. 359-363 vol. 1.

[3] W. G. Homeyer, “Advanced power converters for More Electric Aircraft applications,” in Proc. of Energy Conversion Engineering Conference, 1997, pp. 591-596 vol. 1.

[4] H. Ohashi, “Power electronics innovation with next generation advanced power devices,” in Proc. of INTELEC, 2003, pp. 9-13.

[5] R.G. Hoft, Semiconductor Power Electronics, New York: Van Nostrand Reihold Company, 1986.

[6] D.W. Hart, Introduction to Power Electronics, Upper Saddle River, NJ: Prentice Hall, 1997.

[7] N. Mohan, T.M. Undeland, W.P. Robbins, Power Electronics, 2nd Edition. New York: John Wiley & Sons, 1995.

[8] C.A. Canesin, “Comparison of experimental losses among six different topologies

for a 1.6 kW boost converter, using IGBTs,” in Proc. of PESC, 1995, pp. 1265-1271 vol. 2.

[9] L. Huber, “A design approach for server power supplies for networking

applications,” in Proc. of APEC, 2000, pp. 1163-1169 vol. 2.

[10] J. Yungtaek, “A new, soft-switched, high-power-factor boost converter with IGBTs,” IEEE trans. on Power Electronics, vol. 17, pp. 469-476, July 2002.

[11] E.G. Cook, “Design and testing of a fast, 50 kV solid-state kicker pulser,” in Proc. of Power Modulator Conf., July 2002, pp. 106-109.

[12] M.P.J. Gaudreau, “Solid-state pulsed power systems for the Next Linear Collider,” in Proc. of Particle Accelerator Conf., 2003, pp. 547-549 vol. 1.

Page 25: Research

10

[13] W. Jiang, “Compact solid-State switched pulsed power and its applications,” Proceeding of the IEEE, vol. 92, No. 7, pp. 1180-1196, July 2004.

[14] M.A. Kempkes, “Crowbar replacement through solid-state opening switches [VED applications],” in Proc. of IVEC, 2004, pp. 271-272

[15] R.J. Barker, High-Power Microwave Sources and Technologies, New York: Institute of Electrical and Electronics Engineers, Inc., 2001.

[16] CPI Gyrotron Primer and Specifications Sheets.

[17] CPI Private Communication with Mr. Parent and Dr. Felch.

[18] R.L. Ives, “Design of a multistage depressed collector system for 1 MW CW gyrotrons. II. System consideration,” IEEE Trans. Plasma Science, vol. ED-27, pp. 503-511, April 1999.

Page 26: Research

11

CHAPTER 2.0 THEORETICAL CASCADE BOOST DESIGN

2.1 Switch-Mode Converters

The subject of switch-mode converters is a significant portion of power

conditioning electronics. The general design concept of switch-mode converters is the

delivery of controlled and regulated power from an unregulated source [1]. These

circuits are used in power processing systems integrated into power supplies, motor

control systems, computers, consumer and military electronics, electric transportation

drives, heating and cooling, and other applications where controlled and regulated power

is necessary [2]. Switch-mode converters are designed to sustain a constant output and

compensate variation from an unregulated power source. The focus of the following will

be on DC-DC switch-mode converters, which are designed for DC input and conditioning

of the DC output voltage. A high-level diagram of a DC-DC switch-mode converter is

shown in Figure 2.1. Here, the input and output diagnostics feed data to the controller.

The switch-mode converter’s diagnostics and control loop are designed to give the

necessary switching parameter compensation to maintain the regulated DC output.

Page 27: Research

12

Switch-mode converters operate by repetitively transferring energy from the input

source to the output load by controlled switching devices, hence the name switch-mode

converter. At steady state, the converter control system will maintain an average energy

transfer to the output by adjusting switching characteristics to account for unregulated

input variation. A controllable switching parameter is the switching duty ratio, which is

the ratio of on-time of the switching device(s) to the total time duration of the switching

period as seen in equation [2.1] below:

on onon

t tD = = tSwitching period Ts

f≡ , [2.1]

where f is the frequency in Hertz, and ton is the time the switch is conducting. The output

of switch-mode converters can be adjusted by altering the duty ratio of the switch, or

switches, which in turn adjusts the average energy transfer through the converter. This

control is called pulse width modulation (PWM).

Certain DC-DC switch-mode converter topologies are specifically designed to

step-up or step-down DC voltage. For instance, the buck converter topology creates a

Figure 2.1 General diagram of a generic DC-DC switch-mode converter.

Switching Converter

Unregulated DC

Regulated DC

Diagnostics

Controller

Load

Page 28: Research

13

less-than-unity voltage gain—it is designed to step-down a DC voltage to a lesser DC

voltage. The boost converter has a greater-than-unity voltage gain—it is designed to

step-up DC voltage. Also, some topologies, such as the buck-boost and the Ćuk

converter, are designed to do both—increase or decrease the DC output voltage

depending on the switching duty ratio. Thus stepping-up or stepping-down the voltage is

analogous to a transformer’s winding ratio in AC signal applications. However, the

controllability of the PWM allows switch-mode converters to have real-time adjustable

voltage ratios. The PWM controlled output compensates from variation in the input

source, or load, while delivering a steady DC voltage. Also, DC output filtering is

implemented in most converters for mitigation of the voltage ripple transferred to the

output from the active switching. For this reason, switch-mode converters are

implemented in applications where constant, regulated output is necessary.

2.2 The Boost Converter

The basic boost converter is comprised of an inductor, a switch, a diode, and a

capacitor for output DC filtering of the load. A schematic of the boost converter is

shown in Figure 2.2. The converter stores magnetic energy in the inductor and then

switches it to the output load, through the diode, with every switching period. The output

capacitor provides low-pass filtering, where the average diode current is delivered to the

load.

Page 29: Research

14

The average diode current through the load provides the output DC voltage, but

there is a certain amount of AC voltage ripple coupled to the output. The amplitude of

the AC voltage ripple is affected by the output capacitor, the impedance of the load, and

the switching frequency of the converter. If the output maximum voltage ripple is

specified to be ± 0.5%, the filter capacitance can be designed to meet this specification.

Figure 2.2 Single Stage Boost Converter

Deriving the gain equation for the boost converter is conditional on the

converter’s operational mode. A boost converter can be operated in two modes:

continuous conduction mode (CCM) and discontinuous conduction mode (DCM). The

current waveform of the inductor indicates which conduction mode the converter is

operating. If the current in the inductor does not reach zero during the time when the

switch is open, the circuit is in CCM. If the boost circuit inductor current does go to

zero, it is in DCM. The inductor current for DCM and CCM is shown in Figure 2.3 and

Figure 2.4, respectively. The voltage gain in DCM and CCM is derived from the

principle of balanced inductor volt-second product, where the average voltage across the

inductor is zero during the steady state operation of the circuit.

Page 30: Research

15

Figure 2.3 Discontinuous Inductor Current. D is the switch duty ratio and ∆1 is the diode duty ratio or the fall

time of the inductor current.

Figure 2.4 Continuous Inductor Current. During the fall time of the inductor current, the current remains greater than zero.

For CCM, the derivation of the voltage gain is as follows. When the switch is

closed, the voltage across the inductor is equal to the supply voltage, Vs, where current

increases linearly through the inductor and switch. When the switch opens, the voltage

across the inductor is equal to the supply voltage minus the clamped voltage across the

capacitor, or output voltage, Vo. The inductor voltage is negative during this interval

(toff) and the inductor current will be transferred to the load, through the diode. The

average voltage, LV , across the inductor is equal to zero, as described in equation [2.2]

below:

[ ]L s s o1V = V DT (V - V )(1 - D)T 0T

+ = , [2.2]

DT toff

iL

IL

T

DT ∆1T

iL

Imax

T

Page 31: Research

16

where D is the duty ratio, T is the period, DT equals the time the switch on-time, and (1-

D)T equals the time the switch off-time. Solving for the ratio of output to input voltage,

using equation [2.2], will result in the following step-up voltage gain equation for CCM

is shown below:

o

s

V 1V (1-D)

=. [2.3]

A similar equation for DCM operation will also be derived. The inductor current

is shown in Figure 2.3, where ∆1T is the duration during which the inductor current falls

from Imax to zero. The time, ∆1T, is also the diode forward conduction time, where ∆1 is

the diode duty ratio. To determine the voltage gain in DCM, the same procedure is

followed as in CCM. The average voltage across the inductor, LV , during each period is

zero as shown in the following:

( )L s s o 11V = V DT+ V -V ∆ T =0T

⎡ ⎤⎣ ⎦ . [2.4]

The voltage gain from equation [2.4] yields the following relationship in equation [2.5]:

o 1

s 1

V (D+∆ )V ∆

= . [2.5]

Page 32: Research

17

The value of the interval ∆1, however, is dependent on the boost converter’s inductor,

switching frequency, switch duty ratio, and the output load as seen in equation [2.6]:

o1

s

V 2LV RDT

⎛ ⎞⎛ ⎞∆ = ⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠

. [2.6]

Inserting equation [2.6] into equation [2.5] yields equation [2.7] for the gain of a boost

converter in DCM:

2

o

s

V 1 2D RT= 1+ 1+V 2 L

⎛ ⎞⎜ ⎟⎜ ⎟⎝ ⎠

. [2.7]

2.3 Discontinuous Mode

The cascade boost converter has been designed using DCM due to the necessary

high voltage boost ratio and operation into a high impedance load. An important

advantage of DCM is that the boost converter can operate with a high step-up gain and

allow for reduced inductor size. Figure 2.5 displays a voltage step-up gain comparison of

a boost converter in DCM and in CCM. This shows a favorable linear relationship of the

gain with the switch duty ratio. Table 2.1 shows the CCM and DCM voltage gain ratios,

in relation to the minimum inductor value for operation in the CCM. The gain value for

both modes is independent of converter voltage input, output, or power specifications.

This shows that the inductor for CCM must be several times larger than the inductor used

Page 33: Research

18

in DCM for an equivalent step-up voltage gain. The inductor can consume a large

portion of the converter volume and mass, where power density of the converter can be

increased if the inductor size is minimized.

Unlike CCM, voltage gain for the discontinuous mode does not exclusively

depend on the switch duty ratio, D. Operating in DCM may have advantages over CCM

under the condition of a high impedance load, where the average output current is small.

For instance, in CCM, the duty ratio must be 90% of the switching period to achieve a

voltage gain of ten. In DCM, this gain can be accomplished with a duty ratio of 60%,

where ∆1 is 6.67% of the switching period. Operating at high duty ratios has

disadvantages, such as limited range for the control system to increase the duty ratio

further in order to maintain a regulated output. Thus, the converter’s ability to maintain a

constant output voltage is relatively limited in the CCM.

Page 34: Research

19

0

2

4

6

8

10

12

14

16

18

20

0 0.2 0.4 0.6 0.8 1Duty Ratio

Vol

tage

Rat

io

DCM

CCM

Figure 2.5 Comparison of DCM to CCM (R=25 kΩ, f=10 kHz, L=3.3 mH)

Table 2.1 Comparison of voltage gain and inductor values over a range of duty ratios and voltage gain ratios.

DCM inductor held at 3.3 mH; CCM Lmin adjusted to keep boost converter in CCM.

Duty Ratio CCM Gain CCM Lmin (mH) DCM Gain0.1 1.11 1098 2.430.2 1.25 976 4.270.3 1.43 854 6.120.4 1.67 732 7.980.5 2.00 610 9.850.6 2.50 488 11.710.7 3.33 366 13.580.8 5.00 244 15.440.9 10.00 122 17.31

0.95 20.00 61 18.240.98 50.00 24.4 18.80

Another important aspect of the DCM is a reduction of the reverse recovery loss

associated with typical power diodes in the boost circuits. The reverse recovery loss

occurs during the transient when the diode switches from conducting current in the

Page 35: Research

20

forward biased state to the reverse bias or off state. The reverse recovery current is

formed as residual junction charge is transferred out of the diode. The amount of reverse

recovery loss depends on the design of the diode, where power diodes and high voltage

diodes tend to recover more slowly than lower voltage rated high-speed diodes. Also, the

rate at which the current falls from conducting to blocking influences the peak reverse

recovery current, where higher dIf/dt generally results in higher peak reverse recovery

current. Manufacturers typically list reverse recovery characteristics from a given current

level and rate of current fall. The reverse recovery current, multiplied by the rising

reverse voltage, will result in power loss. Figure 2.6 shows the diode current in the

CCM. The high diode dif/dt in the CCM occurs as the diode turns to the off state, and the

reverse recovery can become significant. Additionally, the reverse recovery current

increases switch loss at turn-on, since the reverse recovery current will contribute to the

switch turn-on current level in the CCM.

The DCM reduces the peak reverse recovery current since the dIf/dt of the diode

is equal to the dIL/dt of the inductor, which can be orders of magnitude lower. The

diagram in Figure 2.7 of DCM diode and inductor currents shows the inductor-limited

dIf/dt diode current. Reduction of the reverse recovery loss can be lessened further by

use of new diode technology. Recent manufacturers such as Cree Inc. have developed

Silicon Carbide (SiC) Schottky diodes with 1200 V and 10 A ratings, that reduce reverse

recovery loss dramatically. Schottky diodes have no minority carrier stored charge, as it

is a majority carrier device, and with SiC, these diodes can have voltage ratings nearly

ten times that of Si Schottky diodes.

Page 36: Research

21

Figure 2.6 The CCM diode and inductor currents. High dif/dt occurring at diode turn-off.

Figure 2.7 The DCM diode and inductor currents

2.4 Cascade Boost Converter Design

The following section details the principles and the derivation of the system

equations used to determine component values under specified switching parameters,

voltage ripple, and load. Figure 2.8 below is an idealized schematic of the two stage

cascade boost converter. The analysis will determine L1, L2, C1, and C2 in an algebraic

form.

dID/dt = IL/dt

Diode current Inductor current

i(t)

High dif/dt

Diode current Inductor current

i(t)

0

IL(t) ID(t)

Page 37: Research

22

R L

D 2 L 2 L 1 D 1

C 2 S 2 C 1 S 1 V s

Figure 2.8 Simplified Cascade boost converter schematic.

2.4.1 Voltage Gain and Inductor Derivation

The cascade boost converter transfers energy through two series boost converter

stages where the output voltage of the second stage is maintained across the load, RL.

The resultant converter voltage gain is the product of the first stage voltage gain and

second stage voltage gain. This voltage multiplication is achieved without use of a

transformer. Each stage operates just as any boost converter, except the first stage does

not operate into a resistive load; rather the stages are coupled together by a controlled

average current flow through the stages. The first stage average output current is equal to

the average input current of the second stage. The effective load impedance for the first

stage will change if the average current through L2 increases or decreases from the PWM

controlled second stage. Thus, the first stage switching duty ratio must be controlled to

compensate for any changes in the average current drawn from the second stage.

The energy transfer process during steady state operation is as follows. During

time, DT or ton, of the switches, a linear-rate rise in current is drawn through L1 from the

DC input source, and L2 similarly draws current out of C1. An example of the inductor

current waveforms is illustrated in Figure 2.9. When both switches are open (time

Page 38: Research

23

duration T-DT), current through diode D1 flows through C1, and simultaneously, current

flows through diode D2 to the output load, filtered by C2. Also shown in Figure 2.9 is the

current through C1, where the charge flowing out of C1, during DT, must equal the charge

flowing into C1 during ∆2T, to maintain a constant capacitor voltage. This shows that

changes in the average second stage inductor current will affect accumulative charge

storage in C1 and thus first stage DC output voltage.

Figure 2.9 Example of the current waveforms of L1 (red), L2 (blue), and C1 (black). D is the switch duty

ratio, and ∆1 and ∆2 are the duty ratios of the diodes and fall time of the inductor current.

The calculation of the inductance values for inductors L1 and L2 is found by

assuming the circuit is in the steady state mode. As mentioned above, the average current

of the capacitors is zero, and the voltage across them equalizes to a constant DC voltage;

the average rate of charge input equals the average rate of discharge. Therefore the

integral of current over one period is zero:DT T

c c

0 DT

i dt= i dt∫ ∫ , charge flows into the capacitors

∆2T

iC1

iL1 iL2

T

DT

∆1T

i(t)

L1 Current L2 Current C1 Current

Page 39: Research

24

from 0 to DT and out of the capacitors from DT to T. If the net current per period is zero,

the net change in voltage per period is also zero: -∆vc = +∆vc. The governing equation of

the capacitors’ voltage is shown below in equation [2.8]:

c dt

c iQv = =

C C∆

∆ ∫ . [2.8]

The following analysis will find the relationship of L1 and L2 using steady state

voltage equalization of C1, where -∆vc1 = +∆vc1 per period. For simplicity, the analysis

also assumes switches S1 and S2 are switching synchronously, where the duty ratio and

period in the following derivation applies to both switches. However, asynchronous

switching may be desirable and will be discussed in section 2.5.

The illustration of the inductor currents that will be discussed is shown in Figure

2.9. The linear increase of the current through L2, while the switches are closed, has a

slope of VC1/L2 for time duration of DT. Therefore, the integral of current flowing out of

capacitor C1 and through inductor L2 is shown below in equation [2.9]:

1 11

DT 2C C

c 2 20

1 V -V (DT)i dt (DT) - DT2 L 2L

⎛ ⎞ = =⎜ ⎟⎝ ⎠∫ . [2.9]

The negative change in the C1 voltage while the switches are closed, then, is given by:

Page 40: Research

25

11 1

DT 2C

C C 1 2 10

1 V (DT)-∆v = i dt= C 2L C∫ . [2.10]

Using the same process when the switches are open results in a slightly more

complicated solution for +∆iC1 and +∆vC1. This is more complicated because while

charge is being added to C1 by iL1, charge is also being removed from C1 by iL2 during the

time ∆2T. Integrating the current through C1, when the switches are open, yields the

positive change in voltage:

11 1

Td C

C C 1 21 1 1 2DT

1 1 1 V 1 V+∆v = i dt= ∆ T DT- ∆ DTC C 2 L 2 L

⎛ ⎞⎜ ⎟⎝ ⎠∫ . [2.11]

Setting –∆vC1 = +∆vC1 and solving for the average capacitor voltage, VC1, results in the

equation [2.12] below:

12 1

C s1 2

L ∆V = VL D+∆

⎛ ⎞⎜ ⎟⎝ ⎠

. [2.12]

Since the design of the cascaded boost converter’s voltage gain or “boost” in the

first and second stages is calculated in DCM, the voltage gain of both stages is in the

form of equation [2.5]. The first stage voltage gain and second voltage gain are shown

below in equations [2.13] and [2.14]:

( )1 1C

s 1

D+∆V =V ∆

, [2.13]

Page 41: Research

26

2

1 1

C o 2

C C 2

V V (D + ∆ )= =V V ∆

, [2.14]

where the second stage output voltage, VC2 is the cascade boost output voltage, Vo.

The values of ∆1 and ∆2 can be solved from [2.13] and [2.14] respectively, given

the duty ratio value, D, and the voltage gain for both stages. Once ∆1 and ∆2 are known,

the value of L2 is found by using equation [2.6] and ∆2 as in equation [2.15]:

12 2

2 2C

C

V RDTLV

⎛ ⎞⎛ ⎞= ∆⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠

. [2.15]

After solving for L2, L1 is found using equation [2.12], as shown in the following

equation [2.16]:

11 2

2 1

VL LD V

s

C

∆⎛ ⎞= ⎜ ⎟+ ∆⎝ ⎠. [2.16]

This concludes the derivation of L1 and L2 given the output load resistance,

voltage gain per stage, duty ratio, and switching period with synchronous switching. The

derivation assumes ideal circuit parameters, where realistic circuit losses absorb energy

which lowers the gain and the output voltage. Such losses are analyzed in the cascade

boost converter prototype and are explained in the chapters to follow.

Page 42: Research

27

2.4.2 Filter Capacitors

As a design parameter, the capacitor C1 and C2 values must also be derived. The

values of these capacitors must fulfill the system requirements for voltage ripple

regulation. The voltage ripple is the ratio of the varying AC component to the average

DC value, and is commonly expressed in percentage. For system volume and mass

minimization, the capacitance values should not greatly exceed the value necessary for

the specified voltage ripple, where extraneous capacitance will consume extra volume

and mass of the system.

For the capacitor C1, the ratio of peak-to-peak AC voltage to DC voltage is

calculated with equation [2.17]:

C1

1 1

v ( Q/CV VC C

∆ ∆ )= , [2.17]

where VC1 represents the DC voltage value and ∆vC1 is the peak-to-peak magnitude of the

voltage ripple. The charge transfer ratio ∆Q, or the current flowing out of the capacitor

when the switches are closed, is used in the calculation, to simplify the integration. The

charge transfer ∆Q is:

C1C1

2

1 V∆Q ∆i dt (DT) DT2 L

⎛ ⎞= = −⎜ ⎟⎝ ⎠∫ . [2.18]

Page 43: Research

28

The results from equation [2.18] can be used to calculate the ripple in capacitor C1, as

given in equation [2.19]:

2

C1

C1 1 2

∆v (DT)V 2C L

= . [2.19]

The voltage ripple for the output capacitor, C2, is more complicated, but the

procedure is similar. The change in the capacitor charge is found by integrating the

amplitude of current above the average current, Io, flowing through the diode, where the

average diode current is equal to the DC output current, Vo/RL. Equation [2.20] below is

the voltage ripple across C2 and RL:

( )C2 2 o C112 C2 2 o

C2 o C1 2

∆v L I V2C V ∆ T DT IV V V L

− ⎡ ⎤⎛ ⎞⎛ ⎞= − −⎜ ⎟⎜ ⎟⎢ ⎥−⎝ ⎠⎝ ⎠⎣ ⎦, [2.20]

where VC2 is the DC output voltage and ∆vC2 is the peak-to-peak AC voltage.

Therefore, with the voltage ripple solved in terms of design parameters, the

capacitance values can be found. The equations [2.21] and [2.22] below derive the

values of capacitors C1 and C2 in terms of the voltage ripple across C1 and C2,

respectively:

2

1C1

2C1

(DT)C ∆v2 LV

= ; [2.21]

Page 44: Research

29

( )C22 o C11

2 2 oo C1 2

L I VC 2∆v ∆ T DT IV V L

− ⎡ ⎤⎛ ⎞⎛ ⎞= − −⎜ ⎟⎜ ⎟⎢ ⎥−⎝ ⎠⎝ ⎠⎣ ⎦. [2.22]

2.5 Asymmetrical Switching

Switching loss is a crucial parameter for consideration in the design of the cascade

boost converter. The average power loss in the switches is proportional to the switching

frequency from the energy lost per transition. Maximizing the switching frequency per

stage, while operating within the specified power loss, can optimize the inductor and

capacitor values for improved system power density. For instance, asymmetrical

switching allows the first stage to be operated at a higher switching frequency

independently of the second stage to take advantage of typically faster, lower voltage

rated switches of the first stage.

Asymmetrical switching of the first and second stages is not difficult to achieve,

given that the capacitance of C1 will satisfy the specified voltage ripple regulation

requirement. Capacitor C1 will then couple both stages with a steady DC voltage, and

both stages can have independent switching frequencies. The primary concern is to

control the steady state current and the subsequent DC voltage across C1 and C2.

One method for independent switching design is to first approach the circuit as a

symmetrically switching circuit. The inductance values are solved as they are solved in

equations [2.15] and [2.16]. Because the inductor values are inversely proportional to the

Page 45: Research

30

switching frequency, the inductors are scaled proportionally with a change in the

switching frequency in each stage. For instance, if the first stage is to be operated with a

ten times higher switching frequency than the second stage, the inductance values can

initially be solved using a symmetrical switching design with the first stage inductor L1

scaled to one tenth of the initial value. The reduced inductor value creates a

proportionally steeper inductor current slope, and the average current flowing into C1

from the first stage will remain as in the symmetrical switching case since the time

interval, D1T1, when the switch is closed would change in proportion to the change in

frequency. Therefore, the average current flowing into C1, C1I , in equation [2.23] shows

that if D1T1 is reduced by one tenth, L1 must be reduced by the same factor as seen in

equation [2.23] below:

s sC1 1 1 1 1 1 1 1

1 1 1

1 1 V 1 VI = ∆ T D T = ∆ D TT 2 L 2 L

⎛ ⎞⎛ ⎞⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠

. [2.23]

The preceding discussion and relation of the inductance to the switching frequency

per stage allows for the asymmetrical switching between stages. With the ability to

operate the stages independently, maximum switching frequency for each stage can be

accomplished. Maximized switching frequency results in minimized inductor values and

reduced inductor volume and mass.

Page 46: Research

31

2.6 Initial Design Values and Simulation Results of an Idealized Model

The inductor and capacitor values in the cascade boost converter are determined

from the equations in Chapter 2. These values were used to simulate an idealized circuit

model of the cascade boost converter with an output of 5 kV and 1 kW. Also the output

regulation has been simulated for a ±0.5% voltage ripple. The circuit simulation results

verify the circuit equations which were used to derive the circuit component values. The

circuit model was created and simulated using Orcad Pspice. The circuit model and the

associated circuit element values are shown in Figure 2.10.

In order to use the equations in Chapter 2, certain values must be specified. These

values are given below and were applied to equations [2.13] through [2.16]. A judgment

was made for the step-up ratios and switching frequencies of the first and second stage,

based on the load impedances per stage and the limiting peak inductor currents to limit

power loss and magnetic field radiation. Also, the second stage switching frequency is

limited to half of the first stage switching frequency due to higher anticipated switch loss.

The cascade boost efficiency must operate at 70% or higher; it has been assumed that a

converter operating at less than 70% efficiency will be unacceptable for the end user,

since heat removal can become a significant issue for high power converters. The

asymmetrical switching frequencies of the first and second stages were accounted for, as

discussed in section 2.5. Also, the values of L1, L2, C 1, and C2 were determined on the

basis of constant input voltage, constant gain, and constant output voltage. Due to

asymmetrical switching, the capacitance value of C1 was determined empirically by

sweeping the capacitance to attain the specified voltage ripple in the first stage output.

Page 47: Research

32

The following values were used as an initial parameter space to begin circuit modeling of

the converter. In the actual prototype cascade boost converter, the parameters were

modified, which will be discussed in Chapter 5.

1st stage gain = 5

2nd stage gain = 10

1st stage switching frequency = 20 kHz

2nd stage switching frequency = 10 kHz (limited due to switch loss)

Duty ratio of both stages = 0.5

Load resistance = 25 kΩ

± 0.5% or less Voltage ripple on C1 and C2

Efficiency = 70% or higher (specification for experimental circuit only)

The circuit values derived from equations [2.13] through [2.16]:

∆1 = 0.125; ∆2 = .055

L2 = 3.4375 mH (rounded down to 3.4 mH)

L1 = 76.6 µH

C1 = 28 µF

C2 = 0.383 µF(rounded up to 0.4 µF)

Page 48: Research

33

R1

.0001

D1

Vg2

TD = 0

TF = 1nPW = 50usPER = 100us

V1 = 0

TR = 1n

V2 = 1

D2R2

.0001

+-

+

-

S2

S

VON = 1.0VVOFF = 0.0V

R3.001

L1

76.6uH

L2

3.4mH

Vs100v

C20.4u

0

C128u

RL

25000

+-

+

-

S1

S

VON = 1.0VVOFF = 0.0V

Vg1

TD =

TF = 1nPW = 25usPER = 50us

V1 = 0

TR = 1n

V2 = 1

Figure 2.10 Idealized Pspice circuit model using derived component values. Resistors R1, R2, R3 were inserted into the model to avoid convergence problems.

The non-idealities existing in the model are negligibly small and do not affect the

converter output. The switches in the model are ideal, but incorporate 1 mΩ impedances

during the on-state. The resistance values of R1 and R2 were added as resistance for

simulation convergence, and the diodes have an internal forward voltage of 0.7-1.0 V.

The total power loss from the non-idealities is approximately 2 W, leading to a model

efficiency of 99.8%. These low losses can be neglected in the model with the efficiency

nearly ideal. The output voltage waveform generated by Pspice, in Figure 2.11, shows an

average DC voltage of about 5.07 kV with a voltage ripple of less than ± 0.5%. The

output is greater than 5 kV since L2 was rounded down to 3.4 mH from 3.4375 mH,

which slightly increases the peak inductor current and average charge transfer into C1.

The observed peak-to-peak magnitude of the ripple is 47.7 V, which correlates to a

0.94% peak-to-peak voltage ripple result. The ± 0.5% voltage ripple, specified above, is

Page 49: Research

34

fulfilled with a 0.4 µF capacitance for C2, which was rounded up from the calculated

value of 0.38 µF.

Figure 2.12 shows the voltage across capacitor C1 and the current through L1.

The peak-to-peak voltage ripple across C1 is calculated to be about 0.94% with a

capacitance of 28 µF. The voltage ripple meets the specification of ± 0.5% or less.

Figure 2.13 shows the voltage across capacitor C2 and the current through L2. The

current waveforms through L1 and L2 show the ∆1 and ∆2 duty ratio values of

approximately 0.125 and .055 respectively, which are in agreement with the calculated

values found with equations [2.13] and [2.14]. The current waveforms also indicate the

current peak as 33 A in L1 and 7.5 A in L2. Since the peak inductor current corresponds

to the current level which the switches must turn off, these current waveforms can be

used to select the solid state switches, estimate the switch loss and heat sink size, and

design the inductors for the experimental circuit.

Figure 2.11 Simulated output voltage of the second stage/converter. ∆V = 47.7 V, which correlates to a voltage ripple of less than ± 0.5%.

Output Voltage

5

5.02

5.04

5.06

5.08

5.1

5.12

5.14

0 100 200 300 400 500Time ( µs )

Out

put V

olta

ge (

kV )

∆V = 47.7 V

Page 50: Research

35

Figure 2.12 Simulated voltage across C1 and current through L1.

Figure 2.13 Simulated current through L2 with the second stage output voltage across C2.

The simulation of the idealized circuit model verifies the derived equations of

Chapter 2. Losses in a practical circuit model will affect the cascade boost converter

C1Voltage and L1Current

490

495

500

505

510

0 25 50 75 100 125Time ( µs )

C1

Vot

lage

( V

)

-505

10152025

3035

L1 C

urre

nt (

A )

C2 Voltage and L2 Current

5.04

5.05

5.06

5.07

5.08

5.09

5.1

0 25 50 75 100 125

Time ( µs )

C2

Vol

tage

( kV

)

-1012345678

L2 C

urre

nt (

A )

Page 51: Research

36

voltage gain, where increasing the duty ratio of both stages is necessary to overcome such

losses. Increasing the duty ratio, however, incurs higher peak inductor current, and thus,

higher current switch current and resistance loss. The losses to consider in the cascade

boost converter appear to be a result of the switching loss and the inductor loss. Other

losses include diode loss, and series resistive loss in the capacitors and inductors.

However, the idealized circuit model results demonstrate the circuit equations in Chapter

2, and show the interaction of the circuit component values.

Page 52: Research

37

REFERENCES FOR CHAPTER 2.0 [1] N. Mohan, T.M. Undeland, W.P. Robbins, Power Electronics, 2nd Edition. New

York: John Wiley & Sons, 1995.

[2] D.W. Hart, Introduction to Power Electronics, Upper Saddle River, NJ: Prentice Hall, 1997.

Page 53: Research

38

CHAPTER 3.0 CASCADE BOOST CONVERTER PROTOTYPE DESIGN

The cascade boost converter prototype has been designed and built to satisfy a DC

output of 5 kV, into a 25 kΩ load, from a 100 V DC source using cascaded boost circuits,

and maintain a +/- 0.5% output voltage ripple regulation. The design goals are

summarized in Table 3.1. This system has been designed with commercial off-the-shelf

(COTS) components, and is a bench top system for the analysis of circuit parameters

such as switch, diode, and inductor loss. A photograph of the bench top prototype

converter is shown in Figure 3.1. Switch theory and design involved in the cascade boost

converter is discussed in this chapter, where the voltage requirements for the second stage

switch, diode, output capacitance, and inductor must be designed to the converter’s

output voltage of 5 kV. This requires COTS switches, diodes, and capacitors to be

arranged in series, and requires adequate high voltage insulation within the second stage

inductor. The implementation of leading inductor core technology, and stacked series

device designs for the second stage, also facilitates scaling to higher power levels for

future designs. Fortunately, the first stage output of 500 V allows for conventional use of

widely available commercial components in this voltage range. This chapter describes

the cascade boost converter circuit design and experimental setup, and highlights the

switching theory used in the design of the second stage switch.

Page 54: Research

39

Table 3.1 Cascade boost converter prototype design goals.

Voltage Input 100 V Voltage Output 5 kV Power Output 1 kW

Output Voltage Regulation

+/- 0.5%

Figure 3.1 Bench top cascade boost converter prototype.

3.1 IGBT Switching Design Theory

The Insulated Gate Bipolar Transistor is a hybrid of a bipolar-junction transistor

(BJT) and a Metal-Oxide-Semiconductor Field-Effect Transistor (MOSFET), and has

accrued success has a high power solid state switching device for its fast switching, low

conduction loss, and high impedance control requirement. These qualities are favorable

for the cascade boost converter and applications such as SMPS, power converters, and

Page 55: Research

40

motor control. The implementation of IGBTs in the cascade boost converter will be

discussed in sections 3.2-4, where the first and second stage components will be covered.

The equivalent circuit of an IGBT is shown in Figure 3.2. The terminology for

the IGBT terminals is the following: the collector is the anode, the emitter is the cathode,

and the gate is the control terminal. The IGBT can be modeled as a MOSFET controlled

BJT, where the MOSFET front end forms the high-impedance gate input. Similar to the

MOSFET, the applied voltage from gate to emitter above the threshold level turns the

device on, and a voltage below this threshold voltage turns the device off. The low on-

state collector to emitter resistance is inherent to the conducting properties of the BJT

portion of the IGBT. The IGBT, in effect, has favorable qualities of both the MOSFET

and BJT. However, IGBTs with ratings greater than 1 kV are commonly used in

applications of 100 kHz or less, since switching loss tends be too great at higher

frequencies. The MOSFET, with a faster switching speed but higher on-state resistance,

are ideal for low power switching at frequencies in the kilohertz to megahertz range. As

a compromise between the MOSFET and the BJT, the IGBT was selected as the switch

of choice.

The frequency limitation of the IGBT switches can be linked to input and output

mechanisms of switching losses due to the MOSFET and BJT, respectively. In the

cascade boost converter application, the turn-off operation is of primary concern as the

switches turn-on with zero current conduction and turn-off current at the peak inductor

current. Two significant factors that affect the turn-off switching time and turn-off

energy loss of IGBTs will be the focus in the following theoretical IGBT switching

Page 56: Research

41

discussion. These factors included the internal IGBT capacitance of the MOSFET

structure and the slowly decaying current “tail” during turn off due to the BJT portion.

3.1.1 IGBT Capacitance

The switch capacitances commonly noted on IGBT data sheets include the input

capacitance, Cies, the reverse transfer capacitance (Miller capacitance), Cres, and the

output capacitance, Coes. These capacitance values vary with Vce, and capacitance versus

Vce curves also are typically provided on data sheets. The input capacitance may affect

the gate drive requirement, which is made up of Cres plus the gate to emitter capacitance,

Cge (Cies= Cres+ Cge). The output capacitance may affect the output circuit and is equal to

the collector to emitter capacitance, Cce, plus the Miller capacitance, Cres (Coes= Cce+ Cres).

The capacitances of the equivalent circuit are shown in Figure 3.2, along with the drift

region resistance, Rdrift, which will be discussed in section 3.1.2. The charge stored in the

input capacitance is the only charge that is to be transferred during switching of the

MOSFET, and thus contributes to the loss in the IGBT [1]. The turn-off transition

waveforms are illustrated in Figure 3.3. The MOSFET portion governs the first three

time intervals of the turn off transient: the turn-off delay time, td(off), the voltage rise

time, tvr, and the first interval of current during turn-off, tf1 [1].

Page 57: Research

42

Figure 3.2 The IGBT equivalent circuit including, internal MOSFET and BJT, switch capacitance, and

drift region resistance, Rdrift.

Figure 3.3 The IGBT turn-off waveforms in a clamped inductive load. td(off): dealy time; tvr: voltage rise-

time; tf1: 1st current fall-time; tf2: 2nd current fall-time.

The gate drive circuit has a significant effect on the switching behavior of IGBTs

[2-4]. In general, increasing the rate at which charge is extracted from the gate, which is

stored in the Cies, will shorten the MOSFET switching time and lower switching energy

loss [5]. This is because the gate drive circuit is similar to a capacitor

charging/discharging circuit. For example, the gate driver voltage charges or discharges

the IGBT input capacitance with a time constant of RC G iesR Cτ = , where RG is the total

resistance of the drive circuit, and Cies is the IGBT input capacitance. The total input

MOSFET

BJT

tvr tf1

tf2

td(off)

vce(t)

ic(t)

vge(t)

0

0

t

tvth

Gate

Collector

Emitter

Cce

Cge

Ccg

pnpBJT

MOSFET

Rdrift

Cies= Cres+ Cge Coes= Cce+ Cres Cres=Ccg

Page 58: Research

43

capacitance is fixed internal to the IGBT, and the external gate resistor is used to control

the gate charging and discharge rate. Using small external gate resistance reduces the

time constant and increases the peaks of the gate current pulses. Consequently, as the rate

at which the capacitance is charged and discharged is increased, faster turn-on and turn-

off of the IGBT will result, due of the faster switching of the MOSFET portion.

However, a minimum gate impedance must be considered in relation to the IGBT

application. Switching stress on the device affects reliability and lifetime. The magnitude

of current and voltage, along with high dvce/dt and dic/dt rates, can cause damage to the

IGBT. For instance, the gate resistor must be made large enough to prevent IGBT latch-

up from high collector to emitter voltage time-rate of change (dvce/dt), and to provide

noise immunity in the gate circuit [1]. Latch-up occurs when the pnpn regenerative

action results from the parasitic thyristor within the IGBT doping structure, and the gate

can no longer control the device to turn-off. Furthermore, controlling the switch rate with

the gate impedance reduces the effects of stray circuit inductance in series with the IGBT

collector, which may result in overshoot of Vce during turn-off from high -dic/dt and lead

to increased power loss or even device failure. Also, the external resistor will dampen

the ringing from the inductance in the gate circuit that may cause the gate voltage to ring

above the threshold voltage, resulting in spontaneous turn-on. With paralleled devices

that have varying lengths of gate connection, the gate resistance lowers the effect of

asymmetrical gate lead inductances that may cause triggering delay and asymmetrical

current sharing among the devices [6]. The main consideration is that the minimum gate

resistance must be designed based on the trade-off of mitigating switching energy loss

Page 59: Research

44

and the factors of insidious transient effects that can reduce the IGBT reliability and

lifetime.

The input capacitance is an important parameter to consider for high current,

switching applications, where fast turn-off is important to limit turn-off loss. The IGBT

modules with current ratings of approximately 100A or greater use several IGBT chips in

parallel to share current. When chips are paralleled, the input capacitance increases

proportional to number of dies in parallel. This creates a large equivalent input

capacitance of the module to charge and discharge during switching. The input

capacitance along with the corresponding stray inductance from gate wire bonds limits

the IGBTs gate drive bandwidth [6]. Utilizing the device datasheet information for input

capacitance information, the gate driver must be designed for the IGBT input capacitance

to meet the design specifications for switching speed and power loss. Such a gate driver

must have low resistance output drivers and low stray inductance in order to compensate

for high capacitance. The gate drive requirements for a specific device’s input

capacitance must be met, and the charging/discharging rate of the capacitance must be

considered for mitigation of switching loss. However under certain applications, device

reliability and lifetime during switching may be trade-off factors to fast switching.

3.1.2 The Turn-off Transient of IGBTs

The portion of the turn-off switching transition following the charge extraction of

the input capacitance, or the fourth and final time interval in Figure 3.3, is the decay of

Page 60: Research

45

the collector current before the IGBT completely turns off. This time interval exhibits an

exponential-like current decay, and has been commonly termed the current “tail”. The

current tail is a result of the Bipolar Junction Transistor (BJT) portion of the IGBT’s turn-

off process. The BJT portion is a pnp type transistor made of the p-type collector, n-type

drift region, and the p-type body. Figure 3.4 shows the cross-section of a Non-Punch

Through (NPT) type IGBT, where the p+ body and p+ collector are highly concentrated p-

typed doped regions, and the drift region is a lower concentrated n-type doped region.

The depletion region of the NPT IGBT, developed by a reverse voltage in the off-state,

does not extend past or “punch through” the n- drift region. Conversely, the Punch

Through (PT) type IGBT implements a thin, highly concentrated, n-type buffer layer

adjacent to the p+ collector and allows the depletion region to extend through the n- drift

region and stop in the n-type buffer layer. The NPT is a common type of IGBT for high

voltage ratings due to excellent thermal properties such as negative temperature

coefficient of resistivity and high thermal reliability [6]. However, as will be shown in

this section, increasing the voltage ratings of NPT type IGBTs can increase the turn-off

transition time [7]. Therefore synchronously switching multiple low voltage rated IGBTs

in series can lead to faster switching compared to a single high voltage rated IGBT,

which is exemplified later in section 3.4, with the design of the second stage switch.

When the NPT IGBT is reversed biased, the p-n junction, J2, can be modeled as

an abrupt p+n- junction of a reverse biased NPT PiN diode [1]. A reverse biased junction

of J2 is shown in Figure 3.5. The peak electric field, Epeak, occurs at this junction, and the

Page 61: Research

46

field is zero at the depletion width edges. The electric field slope depends upon the

background semiconductor doping concentration as shown in equation [3.1]:

0

( ) B

s

dE x Nqdx ε ε

= , [3.1]

where NB is the background doping concentration of a semiconductor device, ε0 is the

electric permittivity of free space, and εs is the relative permittivity of the semiconductor.

In the IGBT, electric field expands over the drift region primarily, due to a relatively low

donor doping concentration, ND, compared to the p+ body acceptor doping concentration,

NA. The electric field slope is much steeper in the higher doping concentration region,

and will therefore not penetrate as far into the p+ body, based on equation [3.1]. Also, by

definition, the NPT type IGBT depletion region width never exceeds the drift region

without device breakdown. Therefore, the minimum drift region size, ldrift, is equal to the

depletion region width, Wd, at breakdown (depletion width is maximum).

Page 62: Research

47

Figure 3.4 Cross-section of the NPT IGBT.

Figure 3.5 PNP transistor layers in the IGBT: Epeak occurs at the junction J2, and the depletion region

width, Wd, extends primarily into the n- drift region due to high doping concentration of the p+ body region. The x axis is the distance along the pnp layers, where the zero occurs at junction J2.

Therefore, the drift region thickness ldrift and NB are designed to hold off a

specified reverse voltage. However, when the IGBT is fully on, the resistance of the drift

p+ body

p+ collector n- drift region

Wd J2 J1

E(0)=Epea

x

dE(x)/dx xn xp

E(x)

ldrift

n- drift region

p + body n+ n+

Collector Contact

Gate Contact

Emitter Contact

GateOxide

J2

J1

p + collector

Page 63: Research

48

region becomes a component of the total on-state power loss, where Rdrift, is shown in the

equivalent IGBT circuit in Figure 3.2. The resistance of the drift region is mitigated as

minority carriers are injected from the p+ collector due to the forward biased junction J1.

The diffusion current density of injected free electrons and holes significantly affect the

drift region conductivity, which allows for low forward on-state voltage at high current

density levels. This minority carrier conductivity modulation mechanism is used in

devices such as BJTs and thyristors. When excess charge carrier concentration exceeds

the background doping concentration of a region by several orders of magnitude, a

condition of high level injection exists. Due to the space-charge neutrality condition, the

presence of free electrons is equally high, creating a plasma-like region of free charge

effectively distributed evenly within the drift region.

When the MOSFET portion turns off, the base current into the pnp BJT is

stopped, and the minority carrier injection of holes into the drift region halts. After the

BJT base current is cutoff and J1 is no longer forward biased, the excess carriers that have

been injected from the p+ collector remain stored in the drift region, and must recombine

for the IGBT to fully turn off. The recombination of the carriers within the drift region

drives the current tail at turn-off, which decays at a rate that depends on the carrier

lifetime [6].

The effect of the drift region thickness, and the voltage rating of the IGBT, on the

current tail can be shown using general equations found in semiconductor device theory.

As discussed earlier, the drift region thickness and doping concentration must be

designed to handle the depletion region width at the device’s specified maximum voltage

Page 64: Research

49

application. Using the peak electric field, Epeak, at J2 the depletion region penetration into

the p-body and the drift region can be found. Using Guass’s law and integrating the

charge within the electric field, the peak electric field is given by

0 0

A p D npeak

s s

qN x qN xEε ε ε ε

= = − , [3.2]

where xp is the distance of the depletion region in the p-body, xn is the distance into the

drift region, and εs is the relative permittivity of silicon. The depletion region width, Wd,

is the sum of xp and xn, and again the electric field ramps linearly from the edge of the

depletion region to the peak field, Epeak, as shown in Figure 3.5. By space-charge

neutrality, the charge from both sides of a p-n junction must be equal where,

A p D nqN x qN x= − . [3.3]

Since the doping concentration of the acceptor doped p-body side is much greater than

the n-type drift region in the IGBT, the distribution of the depletion region can again be

simplified as only reaching into the drift region, by equations [3.1] and [3.3]. Therefore,

from equation [3.2], the depletion width Wd in the drift region can be expressed solely in

terms of the donor concentration, ND, as shown in equation [3.4] below:

0 s peakd n

D

EW xqN

ε ε≈ = . [3.4]

Page 65: Research

50

When the maximum voltage is applied across the reverse-biased junction, Epeak

will be equal to the breakdown electric field, Ebr, of the semiconductor material, and the

depletion region will be at the boundary of the drift region and the p+ collector. In

silicon, Ebr = 3x105 V/cm, and the breakdown process occurs when the field becomes

high enough to start impact ionization in the depletion region from high kinetic energy

carriers accelerated in the field, which establishes the process of avalanche breakdown

[8]. The current flowing through the device, after the avalanche process has been

initiated, results in high power dissipation and device failure. The voltage, Vbr, when the

breakdown will occur is found by integrating the electric field distribution over the

depletion region width:

201

2 2s br

br d brD

EV W EqN

ε ε= =

. [3.5]

Also, it has been assumed that the applied voltage at breakdown is much greater than the

contact potential, Φ0, as expressed in equation [3.6] below:

0 2ln A D

i

kT N Nq n

⎛ ⎞Φ = ⎜ ⎟⎝ ⎠

, [3.6]

where ni is the intrinsic carrier density of the semiconductor. When the applied voltage is

at the breakdown voltage, Vbr, the depletion width equal to the width, brdW . The

Page 66: Research

51

thickness of the drift region, ldrift, must be equal to or greater than the depletion region

width at breakdown ( brdW ):

0 s brdrift dbr

D

El WqN

ε ε≥ = . [3.7]

By equations [3.5] and [3.7], brdW is directly proportional to the reverse breakdown

voltage of the device. It is evident that high voltage rated, NPT IGBTs require thick drift

regions and/or increased n-type doping concentration.

During the fully on-state, and at high level injection, the drift region charge

carrier profile and junction J1 can be modeled as a forward biased PiN diode [9]. The

resistivity of the drift region, ρdrift, is inversely proportional to the quantity of free charge

carriers. This relation is shown in equation [3.8]:

1( )

driftn pq n p

ρµ µ

=+

, [3.8]

where µn and µp are the electron and hole mobilities, and where n and p are the free

charge carrier levels of electrons and holes, respectively. At the high level injection

condition, the resistivity is reduced greatly when a high level of excess hole charge

carriers are injected from the p+ collector into the drift region. The high level injection

case requires the concentration profile length of injected holes, called the hole diffusion

Page 67: Research

52

length, Lp, to be longer than the thickness of the drift region [1]. The diffusion length for

holes during high level injection, LHL, is given in equation [3.9] as

HL a HL driftL D lτ= > , [3.9]

where Da is the ambipolar diffusion coefficient and τHL is the hole lifetime during high

level injection. The value of LHL is the average length of time a hole will diffuse before

recombining with the Auger recombination effects, which shortens the lifetime compared

to the low level injection lifetime [1]. The ambipolar diffusion coefficient is defined in

equation [3.10]:

a HL TD Vµ= , [3.10]

where VT is the thermal voltage, and µHL is the hole mobility at high level injection,

accounting for carrier-carrier scattering, which lowers the mobility compared to the low

level injection mobility [1]. The hole diffusion length is stated as the average distance a

hole will diffuse from the junction before it is recombined [8].

However, the distribution of holes in the preceding discussion deviates from the

PiN diode model with the p-n- junction, J2, at the p+ body is involved. At this junction,

the excess carrier concentration must fall to zero since it is reversed biased, which

changes the boundary conditions and the distribution. The carrier profile is similar to the

PiN diode profile until about midway through the drift region, where it begins to fall

towards the junction [10]. In other words, the distribution in the IGBT drift region does

Page 68: Research

53

not have a high level charge injection across the entire width as the PiN diode does. A

simplified distribution of holes without recombination effects appears linear from the

excess concentration at junction J1, down to the excess concentration at junction of J2,

where the J1 and J2 are shown in Figure 3.4 [10]. However, the distribution of excess

hole concentration can be solved by applying the continuity equation for holes with

recombination effects. The boundary conditions of the hole distribution, pn(x), are the

following: pn(x=0) = np∆ and pn(x=d) = 0, where x is zero at J1 and d is the distance to J1

from J2. At steady state, the continuity equation for holes is shown below in terms of the

hole diffusion length and hole concentration is the following:

2

2

2 0n n

HL

p px L

∂− =

∂ [6]. [3.11]

Applying the aforementioned boundary conditions to equation [3.11] and solving the

second-order differential equation, the derived final solution for pn(x), is expressed in

equation [3.12]:

sinh( ) / ( )sinh( / )

n drift HLn

drift HL

p l x Lp xl L

∆ −= , [3.12]

where np∆ is found by applying boundary conditions for the current densities for electrons

and holes: Jn(x=0) = 0, and Jp(x=0) = J [6]. The solution of np∆ , derived with these

boundary conditions, is shown in equation [3.13]:

Page 69: Research

54

tanh2

HLn

p HL

JL dpqD L

⎛ ⎞∆ = ⎜ ⎟⎝ ⎠

[6]. [3.13]

The principle result reveals that the distribution of the injected holes into the drift region

is related to the diffusion length, the current density, and also the thickness of the drift

region. It is clear that if the drift region is made to be larger, the injection of holes must

be made deeper, resulting in more stored charge existing in the drift region to maintain a

low resistance drift region.

When the IGBT begins to turn off, the excess carriers near J1 will only be

removed by the process of recombination. The effect of sweep out does not exist from J1,

where the growing depletion region removes charge carriers in the electric field—the

depletion region spreads from J2. The spreading depletion region from J2 will extend into

the drift region and sweep out some carriers, but the remaining charge carriers that are

not swept out, recombine at the rate depending on the carrier lifetime during high level

injection. If the beginning of the current tail initiates at t = 0, the collector decay

current, ( )tailCi t , is exponential as shown in the following:

( )(0)( ) HL

tail

tBC

HL

Qi t e βτ

τ

= , [3.14]

where QB (0) is the stored charge in the drift region, or pnp transistor base, at the

beginning of the tail, β is the pnp transistor current gain, and τHL is the high level

Page 70: Research

55

injection lifetime [6]. The carrier lifetime is sensitive to temperature where increased

thermal energy in the IGBT can cause current tail elongation [10]. Reducing the lifetime

can expedite the collector current during the tail time according to equation [3.14].

Lifetime reduction techniques such as gold doping, electron irradiation, and localized

proton implantation are used to speed up the recombination process [10]. However,

reducing the carrier lifetime has trade-offs such as lower charge carrier injection levels

and possibly increased resistivity in the drift region, which increases on-state loss.

The latest generation of high voltage IGBTs, such as the Fast Stop or Soft/Light

Punch Through technologies, and have shown reduced losses compared to similar rated

NPT designs [11]. These designs use advanced doping profiles, which add a thin n-type

doped buffer layer adjacent to the heavily doped p+ collector terminal (next to J1). This

buffer layer provides a desired low forward on-state voltage during conduction and

allows for a thinner drift region and faster recombination of carriers to mitigate switching

loss, especially concerning the current tail loss [12]. Also, new trench-gate IGBT designs

have been shown to lower on-state voltage compared to planar DMOS gate structures by

reducing the on-state resistance in the MOSFET portion [13]. These technologies are

employed in recent 6.5 kV rated and high speed COTS, IGBT modules, designed for

rugged industrial and commercial applications.

Page 71: Research

56

3.1.3 IGBT Switching Theory Conclusion

The IGBT switching frequency is partially limited by input capacitance, where the

gate drive impedance can limit the response of the MOSFET turn-off. The gate drive

circuit must be designed for fast discharge of the IGBT’s input capacitance for mitigated

switching time. Furthermore, the external gate drive resistor can be adjusted to

customize the IGBT switching rate if necessary, by adjustment of the gate drive circuit’s

capacitive time constant. Additionally, the turn-off transition of high voltage IGBTs is

relatively slow due to prolonged carrier recombination in the drift region. The drift

region, also the base of the pnp transistor, is typically designed with a certain thickness

and doping concentration for the IGBT’s off-state voltage rating. Since the total turn-off

time will generally increase as the voltage rating increases, IGBTs with high voltage

ratings can affect system performance such as efficiency and switching frequency in

switch-mode converters. Consideration of the device voltage rating must be taken into

account concerning turn-off loss. As an alternative to high voltage IGBTs, high voltage

switching can be accomplished with a stack of multiple lower voltage rated IGBTs in

series [14]. Isolated gate drivers for each IGBT are required in such an assembly, and

synchronized switching and equal voltage sharing must be ensured for reliable IGBT

stack operation, especially in harsh environments [14]. The maximum number of

switches in series, and parallel, is determined by the break-even point where reliability,

complexity, and cost meet the benefit of lowering switching loss relative to using single-

device IGBTs.

Page 72: Research

57

3.2 First Stage Detail

The first stage of the cascade boost converter experimental setup consists of two

Insulated Gate Bipolar Transistors (IGBTs) switched in parallel, four inductors in series,

two diodes in parallel, and two series capacitors for the DC filtering capacitors between

stages.

Figure 3.6 Schematic of first stage components

The IGBTs used for the first stage are International Rectifier IRGP30B120KD-E

1200 V, 60 A rated, Non-Punch Through (NPT), IGBTs, each in a TO-247AD package.

These IGBTs, switched in parallel, are driven by a single Powerex M57962L gate driver

which is powered by a single Powerex M57145L-01 DC-DC converter designed

specifically for the M57962L. The gate drive delivers a bipolar signal of +15.8 V and -

8.2 V. The DC-DC converter draws power from a 12-18 V input, which is provided by

0.24 Ω

400V TVS

Inductors

Rg ~21 Ω

Cree

Nippon Chemi-con

85.4 µH

Power Ten P83C-30033 Power Supply

237 µF

237 µF

(2 series capacitors)

Caddell-Burns SiC diodes

HP 8012B Pulse Generator

400V TVS

12.5 VPower Supply

gate driver& DC-DC converter

18V TVS

(4 series inductors)

0.24 Ω

Powerex International Rectifier IGBTs

Page 73: Research

58

an 18 W, 12.5 V power supply plugged into the 120V AC laboratory power. The gate

resistance for both IGBTs is controlled by a single potentiometer set nominally at 21Ω.

The gate drive input signal source is controlled by a Hewlett Packard, 8012B pulse

generator. The power supply delivering the input power is a Power Ten, P83C-30033

power supply that is rated at 10 kW with a 300 V maximum output voltage.

The development of the first stage involved determination of switching loss for

the IGBT cooling requirements. The first stage switching loss was determined based

upon initial IGBT testing using the circuit shown in Figure 3.6, but operating as a single

stage converter into a 249 Ω resistive load. The resistive load was used to simulate the

average current draw of the second stage. The test converter produced 550 VDC and the

turn-off current level of each IGBT was 25 A. The IGBT loss was determined to be

approximately 40-50 W when implemented in the final cascade boost converter

prototype.

In order to limit the operating temperature below the rated 125º C junction

temperature, each IGBT was clamped to a forced air cooled aluminum heat sink. The

heat sinks are E3045 profiles by Thermaflo, Inc. shown in Figure 3.7, where each heat

sink was cut to 5 in. in length to give a total surface area of 113 in2 per heat sink. The

heat sink air flow is provided by a 24 V fan that provides about 200 CFM. The thermal

resistance for each heat sink is estimated to be less than 0.5 ºC/W. The combined volume

is 40 in3 for the heat sinks. The total volume consumed by both heat sinks in the

prototype setup, including separation between heat sinks and brackets for mounting, is

given in Table 3.3.

Page 74: Research

59

Figure 3.7 Thermaflo E3045 heat sink profile.

The diodes shown in the schematic in Figure 3.6 are two Cree Inc., CSD10120

silicon carbide (SiC) Schottky diodes with a 1200 V, 10A rating each in a TO-247-3

package. These diodes represent an emerging power semiconductor technology using

SiC, which includes a number of advantages over silicon. Such advantages are the

following: approximately 10 times higher voltage blocking ability, higher operating

temperatures, and higher thermal conductivity. The SiC higher electric field strength

enables the production of Schottky type diodes with a 1200 V rating, where Si Schottky

diodes are rarely rated above 200 V. Schottky diodes are advantageous due to their

inherent fast switching speed and low reverse recovery loss compared to Si, 1200 V, p-n

junction diodes. The power loss of Schottky diodes is almost exclusively due to the on-

state conduction loss. The low power loss and low device temperature rise allows the use

of a small heatsink, where both diodes are mounted on a single heatsink (Wakefield 637-

20ABP) that has a rated temperature rise of 55º C with 6 W dissipated from both devices.

The first stage inductor, L1, is a series arrangement of four inductors by Caddell-

Burns. Two of these inductors are rated at 15 µH and two are rated at 22 µH. The

measured inductance showed that the actual inductance of the four inductors in series is

Page 75: Research

60

85.4 µH at 1 kHz, and was measured with a Sencore LC102, Auto-Z impedance meter.

The manufacturer’s product data of the inductors is given in Table 3.2. The “incremental

current” in Table 3.2, was used to define the peak current rating of the inductors, which

represents the current level at which the inductance decreases by 5% from the initial

(zero-DC) value at steady state conditions. The actual inductance of 85.4 µH changes

the duty ratio from 0.5 to 0.523 as specified in section 2.6 with a value of 76.6 µH for L1.

The inductors were selected with an expected peak current of 50A or less; actual

prototype inductor currents are analyze in Chapter 5.

Table 3.2 Caddell-Burns inductor specifications. Incremental current is the minimum current at which the inductance will be decreased by 5% from the initial (zero-DC) value.

Inductance Model # Max. DC

resistance, Ohms

Rated IDC, Amps

Incremental Current, Amps

15 µH 7000-03 .005 24 52 22 µH 7000-05 .007 21 41

The capacitors used for C1 are two Nippon Chemi-con 220 µF, 450 V rated

capacitors in series. The tested capacitance of each capacitor is 237 µF with 0.24 Ω

series resistance, which was measured with a Sencore LC102, Auto-Z impedance meter.

The capacitors are placed in parallel with two, P6KE400A, transient-voltage suppressor

(TVS) diodes in series, which prevents the first stage output from exceeding

approximately 800 V. The gate-to-emitter voltage of the IGBT is also limited to

approximately 18 V with a P6KE18CA, TVS diode, to prevent damage to the IGBT gate

Page 76: Research

61

from spurious voltage spikes; the gate to emitter voltage must not exceed ± 20 V

according to the manufacturer’s recommendations.

The volume and mass of individual components of the second stage are shown in

Table 3.3. This experimental system has the potential for further volume and mass

optimization with redesign of the first stage IGBT cooling. The heat sinks in the first

stage were utilized primarily due to availability and cost, but the volume and mass of the

cooling system may be reduced by a factor of two with a more efficient COTS heat sink

extrusion profile. Also, technology such as the recent developments in micro-channel

heat sinks with flowing, two-phase liquids for server CPUs, VLSI circuits, and power

electronics, may be efficient enough to extract over 500 W/cm2 in the near-term [15-16].

Such technology has the potential to reduce the cooling system volume and mass to a

fraction of the volume and mass shown in Table 3.3. Notwithstanding, the volume and

mass of the prototype cascade boost converter’s first stage provides adequate cooling for

operation of the experimental circuit, which gives a baseline understanding and allows

for projection of the system volume scaled to higher power levels, as is discussed in

Chapter 6.

Table 3.3 Volume and mass of first stage components using COTS components.

Volume: in3(cc) Mass: lbs(g) IGBTs + heat sinks 90 (1,475) 2.13 (966) Gate drive 6.25 (102.4) 0.046 (21) Diodes +heat sink 1.375 (22.5) 0.077 (34.68) Inductors 13.5 (221.2) 1.085 (492) Capacitors 10.125 (16.4) 0.82 (372) Total 121.25(1,987) 4.158 (1,886)

Page 77: Research

62

3.3 Second Stage Detail

The second stage of the cascade boost converter prototype consists of a stack of

six IGBTs in series, a single inductor, six diodes in series, and two output filter capacitors

in series. This section presents the second stage components and includes the design of

the output load. Figure 3.8 shows the schematic of the second stage with the components

utilized for the prototype converter.

Figure 3.8 Schematic of second stage components

The IGBT stack was built with six IGBTs in series, and is designed for

optimization of switching efficiency using COTS IGBTs in a customized stack assembly.

The stack is shown in Figure 3.9. This IGBT stack assembly, which includes isolated

gate drive circuitry and switch over-voltage protection, was custom built by Loree

Diodes: Inductor (Stangenes nanocrystalline core)

12.5 V power supply

25 kΩ load (10 series Ohmite 2K5L225J)

(6 series 10ETF12FP)

DC filter capacitors (2 series 3.5 kV, 5 µF CSI Capacitors)

3.5 mH

IGBT stack: (6 series IRGP30B120KD-E)

Gate Driver

Gate Driver5 µF

Optically isolated gate signals & isolation transformer

5 µF

First stage DC output voltage

+

-

Page 78: Research

63

Engineering to UMC specification for the cascade boost converter. The design of the

stack uses International Rectifier, IRGP30B120KD-E 1200 V, 60 A rated Non-Punch

Through (NPT) IGBT devices. These IGBTs are the same devices as the first stage

IGBTs. The circuit schematics with details of the stack’s isolated trigger board, isolated

power oscillator, and gate drive boards are given in Appendix A. Each IGBT gate signal

is transmitted through fiber-optics to the individual isolated gate driver boards on which

the IGBT is connected. The gate driver boards provide +15 V for turn-on, and 0 V for

turn-off through a 4.7 Ω gate resistor in series with the gate. The heat sink for each IGBT

is a flat milled aluminum plate (measuring 5.15” x 3” x 0.1875”), with 0.5” radius

rounded corners and 3/32” rounded edges. The stack requires a 12 V DC power input,

the trigger signal, and was forced-air cooled at full power operation.

Page 79: Research

64

Figure 3.9 The IGBT stack with six IGBTs in series, operating as the second stage switch.

Six diodes in series form the second stage 5 kV boost diode. The diode stack

consists of International Rectifier 10ETF12FP diodes, each rated at 1200 V and 10 A

average. These diodes come in a TO-220 “FullPack” package, where the collector

terminal is not exposed on the backside of the diode, as is common on TO-220 packages.

No heat sinks are utilized for the diodes since the expected power dissipation is nearly

negligible.

The second stage inductor, L2, was built by Stangenes Industries to the given

inductor specifications of the cascade boost converter of 3.4 mH. The tested inductance

was 3.5 mH at 1 kHz with a Sencore LC102, Auto-Z impedance meter, and is built with

Page 80: Research

65

four nanocrystalline “C” cores with a 3 mm gap spacing. The design expertise of

Stangenes provided a high quality nanocrystalline-core inductor for the design, which is

used as a basis to scale the cascade boost converter to higher power levels in Chapter 6.

The nanocrystalline core material provides the low core loss of ferrite cores, and

combines the high maximum flux density of iron-based cores. Analogous to SiC in the

semiconductor industry, the nanocrystalline core material represents modern technology

for inductor and transformer magnetic cores in power electronics [17-18]. The

inductance of 3.5 mH for L2 results in the duty ratio, D2, to increase from the ideal value

of 0.5 calculated using the equations in Chapter 2 to the actual value of 0.502.

Page 81: Research

66

The load for the cascade boost converter has an impedance of 25 kΩ. To achieve

operation at a full output voltage of 5 kV and 1 kW, ten 2.5 kΩ, 225 W rated resistors

were arranged in series. The load is shown in Figure 3.10. The resistors are Ohmite,

wire wound, vitreous enamel power resistors (2k5L225J). These resistors are rated for 3

kV dielectric voltage hold off, where each resistor was de-rated to handle 500 V and

dissipate only 100 W. A fan provided a constant air flow across all ten resistors, and the

laboratory temperature was also maintained below 25ºC. The load was measured and

found to be 25 kΩ, both before and immediately after a dissipation power level of 1 kW

output. The volume and mass of the individual second stage components is shown in

Table 3.4. As with the first stage, these volume and mass values for the cascade boost

converter prototype will be used for scaling to higher power levels. The load is not

included in the volume and mass measurements.

Page 82: Research

67

Figure 3.10 The output Load with 10 series Ohmite, wire wound resistors for dissipation of the 1 kW output of the prototype.

Table 3.4 Volume and mass of the second stage components.

Volume in3(cc) Mass lbs(g) IGBT stack 177(2,900) 3.35(1,518) Diodes 0.656(10.75) 0.027(12) Inductor 83.125(1,362) 4.76(2,158) Capacitors 77.3(1267) 2.9(1,330) Total 338.1 (5,540.5) 11.04(5,008)

The total bench top cascade boost converter volume and mass is shown in Table

3.5. The total mass of the system is 15.2 lbs, which represents about 15 lbs/kW. This

figure is not lightweight, but reduction in weight is possible with research specifically

aimed at engineering an advanced cooling system and utilizing next generation light-

Page 83: Research

68

weight, high density capacitors, inductors, and switches. The weight and volume of the

prototype is dominated by commercial off the shelf components, and can be substantially

reduced in an aerospace rated system. The total volume of the system is 459.35 in3 or

0.008 m3, which does not include a packaging factor.

Table 3.5 Total cascade boost converter prototype volume and mass.

Volume in3(cc) Mass lbs(g) First Stage 121.25 (1,987) 4.158 (1,886) Second Stage 338.1 (5,540.5) 11.04 (5,008) Total 459.35 (7,527.4) 15.2 (6,894)

3.4 Second Stage Switch Design Optimization

The original design of the second stage switch utilized two high current, 3.3 kV

IGBT modules in series, which resulted in high average switching loss of over 1.1 kW.

This high switching loss limited the total cascade boost circuit efficiency to

approximately 37%. Using the IGBT theory discussed in section 3.3, the six-series stack

of 1200 V IGBTs was designed to replace the two 3.3 kV IGBTs and provide lower

switching loss. Circuit measurement waveforms and improvement of the switching

efficiency performance of the six-series IGBT stack in contrast to the two 3.3 kV IGBT

modules is presented in the following. This experimental evidence validates the

preceding IGBT switching theory in section 3.1.

The 3.3 kV IGBTs are 200 A rated, NPT (Non-Punch Through) IGBTs from

Dynex Semiconductor. Two of these IGBTs are packaged in a single module unit

Page 84: Research

69

designed for half-bridge circuits for application in motor control or the traction industry.

The IGBTs devices in the module package were connected in series, and synchronized by

simultaneous gate signals. Both gate signals operated with 4.4 Ω external gate resistors.

The module has a 6 kV isolated base plate for mounting to a grounded heat sink. The

heat sink used for thermal management is a liquid cooled heat sink from Aavid

Thermalloy (part number: 416101U00000) and was applied with approximately 0.7 gpm

water flow rate. At this flow rate, the heat sink thermal resistance is rated as 0.013 ºC/W.

The total thermal resistance for each IGBT becomes 0.077 ºC/W, and the combined

IGBT maximum power dissipation handling results in approximately 2.6 kW, with the

maximum IGBT junction temperature at 125 ºC.

The performance of the two 3.3 kV and the six 1200 V IGBTs in series

demonstrates a significant difference in switching loss. Figure 3.11 shows the turn-off

voltage and current waveforms of the two 3.3 kV IGBTs in series as the second stage

switch in the cascade boost converter operating with an output of 5 kV and 1 kW. The

second stage schematic is similar to the circuit shown in Figure 3.8, except that two 3.3

kV IGBTs in series replaced the six 1200 V IGBTs. Figure 3.12 displays the power loss

during turn-off, and the integration of this power loss results in a total of 126.5 mJ of

energy loss per pulse. At a switching frequency of 9 kHz, this turn-off energy loss per

pulse equals an average power loss of 1139 W for the two 3.3 kV IGBTs in series.

Page 85: Research

70

2 Series 3.3 kV Dynex IGBT Turn-0ff Waveform

0

1000

2000

3000

4000

5000

6000

8 9 10 11 12 13 14 15 16 17 18

Time ( µs )

Col

lect

or-E

mitt

er V

olta

ge (

V )

0

2

4

6

8

10

12

Sw

itch

Cur

rent

( A

)Second stage switch voltage

Second stage switch current

Current tail

Figure 3.11 Voltage and current waveforms of two 3.3 kV IGBTs in series as the second stage switch of

the cascade boost converter with and output of 5 kV and 1 kW.

2 Series 3.3 kV Dynex IGBT Turn-0ff Loss

0

10

20

30

40

50

60

8 9 10 11 12 13 14 15 16 17 18Tim e ( µs )

Pow

er (

kW )

0

0.04

0.08

0.12

0.16

0.2

0.24

Ene

rgy

(J)

Second s tage switchturn-off lossEnergy Loss

Eof f = 126.5 mJ

Figure 3.12 Power loss waveform of the two 3.3 kV IGBTs in series. The integrated energy loss is 126.5

mJ per pulse, which results in an average power loss of 1139 W.

A comparison of the two IGBT turn-off loss figures is shown in Table 3.6. Table

3.7 shows the necessary switching duty ratio increase to compensate the switching loss in

Page 86: Research

71

the 5 kV output converter. The increased duty ratio increased the peak inductor charging

current and the average input current to the converter from the source power supply. The

duty ratio for the second stage was limited to 0.66 as the current limit and near-saturation

effects of L2 were reached at marginally higher duty ratios. The first stage IGBT loss

also increases with the 3.3 kV IGBTs implemented in the second stage since the first

stage peak switch current must increase proportionally to regulate the voltage across C1.

Due to the maximum first stage IGBT power loss limitations and the second stage

inductor current limit, the switching duty ratios of both stages were maximized in order

to optimize the converter’s step-up ratio. The optimized cascade boost converter’s

maximum voltage gain was only 46.7 with two series 3.3 kV IGBTs for S2. The 5 kV

and 1 kW output was achieved, using an input voltage of 107 V as shown below in Table

3.7. Again, the loss of the 3.3 kV devices resulted in a cascade boost converter efficiency

of approximately 37%.

Table 3.6 Turn-off loss comparison of the two IGBT designs.

Turn-off energy loss

Average turn-off power loss

2 series 3.3 kV, 200 A IGBTs 130 mJ 1139 W 6 series 1200 V, 60 A IGBTs 11.6 mJ 93.55 W

Page 87: Research

72

Table 3.7 Cascade boost converter parameter comparison of 3.3 kV and 1200 V IGBTs.

2 series 3.3 kV, 200 A IGBTs

6 series 1200 V, 60 A IGBTs

Efficiency 37 % 76.6 % Input voltage 107 V 100 V Average input current 24.8 A 13.1 A First stage output 589 V 503 V D1 0.8 .614 D2 .66 0.513 iL1 max 50 A 34.5 A iL2 max 11.3 A 8.1 A

The low system efficiency using 3.3 kV rated IGBTs motivated the redesign of S2

to achieve a lower switching loss. The waveforms of Figure 3.11 and Figure 3.12

demonstrate the theoretical relation between voltage ratings and current tail magnitudes

in IGBTs, where the high voltage rated IGBTs exhibit a long and substantial current tail.

The redesign of S2 with six series 1200 V rated IGBTs experimentally demonstrates a

lower current tail loss and faster switching speed, which results in nearly 92% lower loss

in S2 and increases the efficiency to 76.6% from 37%. Analysis of the IGBT stack turn-

off switching loss and further experimental results of the prototype cascade boost

converter are detailed in Chapter 5.

Page 88: Research

73

REFERENCE FOR CHAPTER 3.0 [1] N. Mohan, Power Electronics, 2nd Edition. New York: John Wiley & Sons, 1995.

[2] A. Raciti, “Control of the switching transients of IGBT series strings by high-performance drive units,” IEEE Trans. on Industrial Electronics, vol. 48, pp. 482-490, June 2001.

[3] Y.C , Gerstenmaier, “Switching behaviour of high voltage IGBTs and its dependence on gate-drive,” in 1997 Proc. ISPSD, pp. 105-108.

[4] R, Chokhawala, “Gate drive considerations for IGBT modules,” 1992 Conf. record on Industry Applications Society Annual Meeting, vol. 1, pp. 1186-1195.

[5] Dynex Semiconductor application note; AN4504 (http://www.dynexsemi.com)

[6] V. K. Khanna, IGBT Theory and Design, Piscataway, NJ: IEEE Press, 2003.

[7] A. Petterteig, J. Lode, and T.M. Undeland, “IGBT turn-off losses for hard switching and with capacitive snubbers,” Conf. Record of Industry Applications Society Annual Meeting, 1991, pp.1042-1049.

[8] B. G. Streetman, Solid State Electronic Devices, 2nd ed., N. Holonyak, Jr., Ed. Englewood Cliffs, NJ: Prentice-Hall Inc., 1980.

[9] S. Lefebvre and F. Miserey, “Analysis of CIC NPT IGBT's turn-off operations for high switching current level,” IEEE Trans. on Electronic Devices, vol. 46, pp. 1042-1049, May 1999.

[10] H.-W. Kim, “Excess carrier density and forward voltage drop in trench insulated gate bipolar transistor (TIGBT),” in 2002 Proc. MIEL Conf., vol. 1, pp. 12-15.

[11] H, Husken, “FieldStop IGBT with MOS-like (tailless) turn-off,” in 2003 Proc. ISPSD, pp. 338-340.

[12] X, Kang, “Characterization and modeling of high-voltage field-stop IGBTs,” 2002 Conf. record on Industry Applications Society Annual Meeting, vol. 3, pp. 2175-2181.

[13] J. Yamada, “Low Turn-off Switching Energy 1200V IGBT Module,” in 2002 Proc. Industry Applications Conf., pp. 2165-2169.

Page 89: Research

74

[14] P.R. Palmer, “A comparison of IGBT technologies for use in the series connection,” in 1996 Proc. Power Electronics and Variable Speed Drives Int. Conf., pp. 236-241.

[15] D. Faulkner, “Practical design of a 1000 W/cm2 cooling system,” in Proc. 19th Annu. IEEE Semiconductor Thermal Measurement and Management Symposium, 2003, pp. 223 – 230.

[16] G. Upadhya, “Closed-loop cooling technologies for microprocessors,” in Tech Digest IEEE Int. Electron Devices Meeting, 2003, pp. 32.4.1 - 32.4.4

[17] M. Ferch, “Light transformers for kilowatt SMPS based on nanocrystalline soft magnetic cores,” in Proc. of Seventh Int. Conf. on Power Electronics and Variable Speed Drives, 1998, pp. 411-415.

[18] W.A. Rease, “Design technology of high-voltage multi-megawatt polyphase resonant converter modulators” in Proc. of 29th Annu. Conf. IEEE Industrial Electronics Society, 2003, pp. 96-101.

Page 90: Research

75

CHAPTER 4.0 DIAGNOSTICS OF THE CASCADE BOOST CONVERTER PROTOTYPE

Evaluation of the cascade boost converter operation requires utilization of the

current and voltage waveforms in the system. The waveform diagnostics are paramount

for gauging component loss and verification of system performance. The voltage

measurements in the first and second stage require accurate voltage diagnostics for

measuring fast-changing switch voltage. Additionally, the second stage of the cascade

boost converter requires voltage diagnostics designed to measure the high voltage DC

and time-varying voltage transients. Also, the current measurements require diagnostics

capable of accurately measuring the current magnitudes in the system. This chapter

describes the diagnostic equipment used in the measurements of the experimental cascade

boost converter.

4.1 Voltage Diagnostics

The voltage diagnostics include DC voltage measurements and time varying

voltage measurements. The attenuation and frequency bandwidth response of the voltage

probes used for the cascade boost diagnostics are shown in Table 4.1. The output DC

voltage of the cascade boost converter was measured with a Fluke 80k-40, 1000X high

voltage probe. The Fluke 80k-40 probe has a maximum voltage rating of 40 kV DC, and

is designed to interface with digital multi-meters with 10 MΩ ( ± 1%) input impedance or

greater. The Fluke 80k-40 probe is connected to a Fluke 23 III, digital multi-meter. For

Page 91: Research

76

the first stage output DC voltage measurement, a Fluke 77 II, digital multi-meter was

connected directly across capacitor C1. The Fluke 23 III and 77 II DC voltage

measurement accuracy is ± 0.3%. The input DC voltage was monitored using the voltage

monitor on the front panel of the PowerTen P83C-30033, input power supply, and the

accuracy has been verified with a Fluke digital multi-meter. These DC voltage

diagnostics provide a convenient method to observe the cascade boost converter’s input,

first stage, and output DC voltages.

The cascade boost converter’s voltage waveforms of the first and second stage

IGBTs and the output voltage ripple were measured with a Tektronix P6015A, 1000X

high voltage probe. The P6015A has a 20 kV maximum DC voltage rating, 75 MHz

bandwidth, and includes a compensation box for probe frequency response

characteristics. A Tektronix P5100, 100X voltage probe, which has a 2.5 kV maximum

rating, was also used to measure the first stage output voltage for measurement of first

stage voltage ripple. The measurement of the IGBT gate drive circuitry was

accomplished with a Hewlett Packard 10071A, 10X voltage probe, which monitored the

switch duty ratio and the switching frequency for each stage of the cascade boost

converter. A summary of the voltage diagnostics is displayed in Table 4.2 below.

Table 4.1 Voltage probes for circuit diagnostics including attenuation factor and bandwidth.

Voltage Probe Attenuation Bandwidth Fluke 80k-40 1000X DC-60 Hz Tektronix P6015A 1000X 75 MHz Tektronix P5100 100X 250 MHz Hewlett Packard 10071A 10X 150 MHz

Page 92: Research

77

Table 4.2 Measurement and diagnostic equipment implemented.

Measurement Diagnostic Equipment Input Voltage Power supply (PowerTen P83C-30033) voltage monitor display First Stage Output DC Voltage Direct connection to a Fluke 23III DMM Second Stage Output DC Voltage Fluke 80k-40 1000X HV probe to a Fluke 77II DMM First Stage Output Waveform Tektronix P5100 100X probe Second Stage Output Waveform Tektronix P6015A 1000X HV probe First Stage IGBT Waveform Tektronix P6015A 1000X HV probe Second Stage IGBT Waveform Tektronix P6015A 1000X HV probe

4.2 Current Diagnostics

All of the current measurements were made using two current monitors. One of

the current monitors is a Pearson Electronics current monitor model 110A. This current

monitor gives an output of 0.1 V/A of current flowing through the conductor being

measured. It has a rating of 10 kA, 0.8% droop/ms, usable rise time of 20 ns, and

maximum RMS current of 65 A. The probe has a 4 inch outer diameter and a 2.1 inner

diameter. These specifications allow the 110A current monitor to accurately measure the

system current.

The second current monitor is made by Stangenes Industries, model SI 17385.

This probe is smaller in size, consuming 4.25 inch outer diameter and 1 inch inner

diameter. The Stangenes current monitor gives an output of 0.1 V/A, has a rating of 5

kA, 1.0 % droop/ms, a usable rise time of 20 ns, and maximum RMS current of 30 A.

These specifications will also satisfy the diagnostic requirements for accurate current

measurement. A summary of the current diagnostics are displayed in Table 4.3 below.

Page 93: Research

78

Table 4.3 Current monitors for circuit diagnostics

Current Monitor Output Peak Current Maximum Continuous Current

Usable rise time

Droop

Pearson 110A 0.1 V/A 10 kA 65 A RMS 20 ns 0.8 %/ms Stangenes SI 17385 0.1 V/A 5 kA 30 A RMS 20 ns 1.0%/ms

4.3 Oscilloscope

The voltage and current measurements utilized a Tektronix TDS3034, four

channel color digital phosphor oscilloscope. The TDS3034 oscilloscope bandwidth is

300 MHz and has a sample rate of up to 5 GS/s. The waveform data was collected and

saved on to a computer disk, where data analysis was completed using Microsoft Excel

and MATLAB. Specific measurement setups for the voltage and current diagnostics and

data analysis processes are discussed further in Chapter 5. The waveforms and

component loss analysis are also presented in Chapter 5.

Page 94: Research

79

CHAPTER 5.0 CASCADE BOOST CONVERTER PROTOTYPE RESULTS

The output, efficiency, and loss analysis of the cascade boost converter prototype,

are described in the following sections, where the prototype converter has satisfied the

goals given in Chapter 3, and are summarized in Table 5.1. The experimental converter

has demonstrated a step-up voltage gain of 50, an output of 5 kV DC at 1 kW, and a

regulation of better than +/- 0.5%. Thus a 100 V DC input was stepped up to 5 kV DC

into the 25 kΩ resistive load, which produces an output power of 1 kW. The gains for the

first and second stages were 5 and 10, respectively. With the switching frequencies

operating at 20 kHz for the first stage and 9 kHz for the second stage, the converter

efficiency resulted in 76.6%.

It is believed that the efficiency improvement is possible with additional

optimization of the circuit, such as lowering the inductor losses and the IGBT switching

loss with faster switches and possibly parallel stages. The detailed analysis of the

component loss is used for extrapolation of the loss per stage, and also enables a realistic

loss estimation of the cascade boost converter scaled to higher output power and voltage

levels.

Table 5.1 Cascade boost converter prototype design goals.

Voltage Input 100 V Voltage Output 5 kV Power Output 1 kW

Output Voltage Regulation

+/- 0.5%

Page 95: Research

80

5.1 Cascade Boost Converter Output and Efficiency

The circuit performance of the converter is evaluated in this section while

operating with a DC output voltage of 5 kV from a 100 V DC input. As discussed in

Chapter 4, the input voltage to the cascade boost converter was measured using the

voltage monitor on the front panel of the input power supply (PowerTen P83C-30033),

the first stage DC output voltage was measured with the Fluke 77 II digital multi-meter.

The second stage output was measured with the Fluke 80k-40, 1000X high voltage DC

probe and Fluke 23 III, digital multi-meter. The switch duty ratios were adjusted

manually to tune the step-up voltage gain to 5 for the first stage and 10 for the second

stage, which resulted in .614 and .513 for the first and second stage duty ratios,

respectively. The output voltage of the first and second stage was also measured with the

Tektronix P6015A, high voltage probe, and the Tektronix TDS3034, oscilloscope. The

first stage output voltage waveform, with an average of 503 V, is shown in Figure 5.1,

and the converter output voltage, with an average output of 5 kV, is displayed in Figure

5.2.

The converter output ripple is measured to be 49.3 V, which corresponds to a

regulation of +/-0.49% and meets the +/-0.5% regulation goal. According to these

results, the 2.5 µF filter capacitance used for the converter appears to be the minimum

capacitance to achieve a +/-0.5% regulation when stray switch inductance of the actual

circuit is present. The stray switch inductance in both stages resulted in a voltage

overshoot across the switches during turn-off, which coupled to the first and second stage

outputs, as can be seen in Figure 5.1 and Figure 5.2, respectively. In order to provide

Page 96: Research

81

regulation within the specified value, the switch overshoot was attenuated by the output

capacitors. In a future design, mitigation of stray inductance from a low-inductance

circuit design can lead to an optimal output regulation with reduced filter capacitance.

Thus the capacitor volume and mass will be reduced and result in a more compact

converter design.

Firs t Stage Output Voltage

400

450

500

550

600

Tim e ( µs )

1st S

tage

Out

put (

V )

Averge voltage = 503 V

Stray switch inductance voltage spike.

∆1T

Figure 5.1 First stage DC output voltage; the average output voltage is displayed at 503 V, for a voltage

gain of 5 from the 100 V input.

Page 97: Research

82

Second Stage Output Voltage

4800

4900

5000

5100

5200

Time ( µs )

2nd S

tage

Out

put (

V ) Averge voltage = 5010 V

∆2T

Figure 5.2 Cascade boost DC converter output voltage; the average output voltage is displayed at 5010 V,

for a voltage gain of approximately 10 from the 503 V first stage output.

The efficiency values were calculated by using the input voltage, the first stage

output voltage, and the converter output voltage. These values were then multiplied by

the average current flow at each stage. The average current values were determined

using the current waveforms measured from the first and second stage inductor currents.

The Pearson 110A and the Stangenes current monitors were utilized to measure these

inductors currents. The first stage inductor current waveform is shown in Figure 5.5, and

the second stage inductor current waveform is shown in Figure 5.6. The time-average

value of each current waveform equals the average input current per stage. Thus, the

time-average power delivered to each stage was calculated as the following: [5.1], [5.2],

[5.3]

Page 98: Research

83

( )( )0 0 0DCP V I= ; [5.1]

( )( )1 1 1DCP V I= ; [5.2]

( )( )2 2 2DCP V I= , [5.3]

where P0 is the average input power, VDC0 is the input voltage, P1 is the average first

stage output power, VDC1 is the first stage output voltage, P2 is the average second stage

output (converter output) power, VDC2 is the converter output voltage, and 0 2I I− are the

average current values from the respective stage.

Table 5.2 summarizes the average values measured in the cascaded boost

converter, where the average current values listed are the time average values of the stage

input currents as discussed above. Table 5.3 summarizes the calculated efficiency and

loss of the system per stage. Also, the input voltage to the cascade boost converter was

gradually increased from 0 V to 100 V, with constant switch duty ratios of .614 and .513

for S1 and S2, respectively. The converter’s first stage output voltage, rising from an

increasing input voltage, are shown in Table 5.4, and the voltages per stage, shown in

Figure 5.3, show a highly linear relationship with the input. The total system efficiency,

first stage efficiency, and second stage efficiency with respect to the output voltage is

shown in Figure 5.4.

Page 99: Research

84

Table 5.2 Average value data of the system operated at 5 kV output.

DC Voltage

Average Current

Average Power

Converter Input 100 V 13.1 A 1,310 W First Stage Output 503 V 2.3 A 1,157 W Converter Output 5,010 V 200.4 mA 1,004 W

Table 5.3 Efficiency and loss for the first and second stages operated at 5 kV output.

Efficiency Power Loss First stage 88.33% 152.9 W Second stage 86.77% 153.08 W Total system 76.64% 305.98 W

Table 5.4 First stage and second stage voltage data of the cascade boost converter prototype as input voltage is varied from 0 V to 100 V.

Input Voltage (V) First Stage Output (V) Second Stage Output (V)

0 0 0 10 48.1 428 20 93.2 931 40 200.3 1,964 50 252.7 2,485 60 300.9 2,984 70 351 3,470 80 401 3,990 90 454 4,530

100 503 5,010

Page 100: Research

85

First & Second Stage Output Voltage

0

1000

2000

3000

4000

5000

6000

0 20 40 60 80 100 120

Input Voltage

Out

put V

olta

ge

Firs t Stage Output

Second Stage Output

Figure 5.3 First stage and second stage output voltages as input voltage is varied from 0 to 100 V, with constant switching duty ratios of .614 and .513 for the first and second stages respectively.

Cascade Boost Efficiency

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

0 1000 2000 3000 4000 5000 6000

Output ( V )

Effi

cien

cy

Total Efficiency

2nd Stage Efficiency

1st Stage Efficiency

Figure 5.4 Efficiency of the individual stages and the overall converter.

Page 101: Research

86

The inductor charging currents of the cascade boost converter prototype shown

below in Figure 5.5 and Figure 5.6 display subtle differences with the inductor currents

produced by the ideal converter simulated and discussed in section 2.6. These

differences are to be expected due to the non-ideal properties and losses of the prototype

converter. One of the obvious differences is the larger duty ratio of the prototype

converter, as compared to the simulation. As the converter losses accrue, the duty ratio

must be increased to store more energy in the inductor that provides compensation for the

losses. The rising current slope of the first stage inductor, indicating a closed position of

the first stage switch, accounts for about 30.1 µs of the total 49 µs switching period. This

corresponds to a 0.614 duty ratio, where the ideal converter operated at a duty ratio 0.5.

In addition to the losses, the larger duty ratios of the prototype converter are a result of

larger inductor values. The actual inductance utilized in the first stage is 85.4 µH as

described in section 3.1, which is larger than the 76.6 µH ideal inductance value. The

larger inductance is evident by the lower time-rate of change, or di/dt, of the rising first

stage inductor current. The same effect is seen in the second stage, where the actual

inductor is 3.5 mH and the ideal converter inductor is 3.3 mH. The duty ratio of the

second stage is 0.513 in the prototype converter, and 0.5 in the ideal converter simulation.

Page 102: Research

87

Firs t Stage Input Current

-505

10152025303540

0 5 10 15 20 25 30 35 40 45 50 55 60Tim e ( µs )

Inpu

t Cur

rent

( A

)Experimental CurrentSimulated Current

Figure 5.5 First stage experimental and simulated input currents when the converter is operated with a 100

V input and 5 kVoutput.

Second Stage Input Current

-2

0

2

4

6

8

10

0 20 40 60 80 100 120

Tim e ( µs )

Inpu

t Cur

rent

( A

)

Experimental Current Simulated Current

Figure 5.6 Second stage experimental and simulated input currents when the converter is operated with a

100 V input and 5 kV output.

Page 103: Research

88

5.2 Component Loss Analysis

The loss analysis of the cascade boost converter prototype operating at 5 kV

output indicates that the IGBT switching loss and inductor loss are the two most

significant components of the total power loss. The following discussion estimates the

switching, diode, and capacitor loss of both stages from analysis, and from these, the loss

in the inductor was derived. The loss analysis also provides a baseline for the loss of a

cascade boost converter scaled to higher power levels. For all of the following

component loss analysis, the cascade boost converter operating conditions are shown

below in Table 5.5.

Table 5.5 Cascade boost converter operating conditions for the component loss analysis

Circuit Conditions Input Voltage 100 V First Stage Output 503 V First Stage IGBT Frequency 20 kHz First Stage Duty Ratio 0.614 Second Stage Output 5010 V Second Stage IGBT Frequency 9 kHz Second Stage Duty Ratio 0.513

5.2.1 Measurement of the First Stage IGBT Loss

The loss of the first stage IGBTs is determined from the measured voltage and

current waveforms during operation of the cascade boost converter at the full 5 kV

output. The Stangenes CT was utilized to measure the combined current of the parallel

diodes comprising diode D1, and the Pearson CT was setup to measure the inductor

Page 104: Research

89

current of L1. The combined current of the parallel IGBTs is found by subtracting the

diode current from the inductor current. This method allows for the measurement of L1,

D1, and S1 current with two current monitors. The voltage probe setup includes the

P5100 probe measuring the voltage across C1, and the P6015A probe across the collector

and emitter of one IGBT. The data recorded by the Tektronix TDS3035 oscilloscope,

was saved and analyzed as a spreadsheet in Microsoft Excel. Utilizing functions in

Excel, the voltage and current waveforms were multiplied together to produce the

waveforms for power loss. The power loss waveforms were integrated, using a custom

routine in MATLAB, for calculation of energy loss per period. Average power loss is

calculated by multiplying the energy loss per period by the switching frequency as in

equation [5.4] below:

/0

( ) ( )T

loss s sloss periodP f E f v t i t dt∗ ∗= = ∫ , [5.4]

where v(t) and i(t) are the instantaneous voltage and current values. The MATLAB

routine that was used for integration is given in Appendix B.

The turn-off loss and conduction loss, during the on-state, were processed

separately and the respective average powers were added together to derive the total

switch loss. The derivation for the average conduction loss of the IGBTs is discussed in

Appendix C. The IGBT turn-on loss was neglected due to the zero turn-on current

inherent in discontinuous conduction mode of the cascade boost converter. Also, the

Page 105: Research

90

slow time-rate of change of switch current, after turn-on, is limited by the inductance of

L1.

Figure 5.7 shows the turn-off voltage and current waveforms of the two first stage

IGBTs in parallel, where the current is sum of both collector currents. The first stage

IGBT turn-off demonstrates decaying current after the initial fast drop in current,

commonly termed the “current tail”. The current tail increases the turn-off loss

significantly and is an inherent characteristic which results from the minority carrier

recombination and the bipolar junction transistor component of the IGBT. According to

IGBT theory, the current tail does not scale linearly with the IGBT current at turn-off; the

current tail loss significance reduces as the turn-off current is increased [1]. This

explains the increasing system efficiency as the input voltage is increased, as seen in

Figure 5.4.

Figure 5.8 shows the IGBT turn-off power loss waveform for the two first stage

IGBTs, where the cascade boost is operating at the conditions shown in Table 5.5.

Average power loss calculated from these waveforms is 49.3 W per IGBT or 98.57 W

total. Table 5.6 displays a summary of the IGBT waveforms and average power loss.

Page 106: Research

91

First Stage Switch Turn-0ff Waveform

0

100

200

300

400

500

600

700

0.5 1 1.5 2 2.5 3

Time ( µs )

Col

lect

or-E

mitt

er V

olta

ge (

V )

0

5

10

15

20

25

30

35

40

Sw

itch

Cur

rent

( A

)

First stage switch voltage

First stage switch current

Current tail

Figure 5.7 Turn-off voltage and collector current waveforms of the first stage IGBTs in parallel.

First Stage Turn-0ff Loss

0.0

2.0

4.0

6.0

8.0

10.0

12.0

14.0

16.0

18.0

20.0

0.5 1 1.5 2 2.5 3Time ( µs )

Pow

er (

kW )

Eof f = 4.83 mJ

Figure 5.8 First stage turn-off power loss. The integrated turn-off loss = 4.83 mJ at 20 kHz for 98.57 W

average loss.

Page 107: Research

92

Table 5.6 Summary of the first stage IGBT data.

Switching period (frequency) 49 µs (20.4 kHz) Duty ratio 0.614 Ic 34.5 A Vce 503 V Voltage rise time (10-90%) 94 ns Current fall time (90-10%) 240 ns Conduction loss per IGBT 12.55 W Turn-off loss per IGBT 36.75 W Total IGBT loss 98.57 W

5.2.2 Measurement of the Second Stage IGBT Loss

The second stage IGBT loss was determined by the same process as the first stage

IGBT loss. The diagnostic equipment used to monitor the voltage and current includes

the Pearson and the Stangenes current transformers, and two Tektronix P6015A 1000X

probes. The Pearson probe measured the total emitter current of the IGBT stack. One of

the P6015A probes measured the collector voltage, and the other was used to measure the

converter output voltage. The turn-off voltage and current waveforms and turn-off loss

are shown in Figure 5.9 and Figure 5.10 below.

Page 108: Research

93

Second Stage Turn-0ff Waveform

0

1000

2000

3000

4000

5000

6000

0 1 2 3 4 5Time ( µs )

Col

lect

or-E

mitt

er V

olta

ge (

V )

0

1

2

3

4

5

6

7

8

9

Em

itter

Cur

rent

( A

)

Second stage switch voltage

Second stage switch current

Figure 5.9 Turn-off voltage and current waveforms of the second stage IGBT stack.

Second Stage Turn-0ff Loss

0

5

10

15

20

25

30

35

40

1 1.5 2 2.5 3 3.5 4Time ( µs )

Pow

er (

kW )

Eof f = 11.6 mJ

Figure 5.10 Second stage IGBT stack turn-off power loss. The integrated turn-off energy loss = 11.6 mJ at

9 kHz for 104.7 W average loss.

Page 109: Research

94

Table 5.7 Summary of second stage IGBT data.

Switching period (frequency) 110.6 µs (9.04 kHz) Duty ratio 0.513 Voltage rise time (10-90%) 586 ns Current fall time (90-10%) 438 ns Current at turn-off 8.1 A Off-state voltage 5010 V Conduction loss per IGBT 2 W Turn-off loss per IGBT 15.22 W Total IGBT stack loss 104.7 W

5.2.3 Measurement of Diode Losses

The diode loss in the first stage was measured, where the diagnostic setup for

measuring the diode current and voltage waveforms is identical to the measurements of

the first stage IGBT waveforms. However, a differential voltage measurement was taken

by subtracting the output voltage from the IGBT collector voltage to obtain the voltage

across the diode. The analysis of determining average power loss is also the same as the

IGBT average power loss. High frequency oscillations and noise occurring in the

following waveforms are significant, but considered to be a result of stray capacitance

and inductance in the diagnostic equipment and circuit. Such oscillations produce zero

net energy loss after integration as they are purely reactive. Figure 5.11 and Figure 5.12

display the diode current and voltage waveforms.

The parallel SiC Schottky diodes exhibit very low reverse recovery current loss.

The small negative current aberration following the 8 µs mark may be due to junction and

stray capacitance as a result of the time rate of change in the diode voltage [1-3]. The

first stage diode average power loss has been determined to be 1.67 W from an integrated

Page 110: Research

95

energy loss of 0.87 mJ per pulse. As a comparison, the diode loss represents only 1.7%

of the IGBT loss. This loss generally agrees with the very small temperature rise of the

heat sink on which the diodes are mounted.

First Stage Diode Waveforms

-100

-80

-60

-40

-20

0

20

40

60

80

100

0 1 2 3 4 5 6 7 8 9Time ( µs )

Dio

de V

olta

ge (

V )

-5

0

5

10

15

20

25

30

35

Dio

de C

urre

nt (

V )

First stage diode voltage

First stage diode current

Figure 5.11 First stage diode voltage and current waveforms.

Page 111: Research

96

Firs t Stage Diode Power Waveform

-500

0

500

1000

1500

2000

0 1 2 3 4 5 6 7 8 9Tim e ( µs )

Pow

er (

W )

Eloss=0.82 mJ

Figure 5.12 First stage diode power loss waveform. Integrated energy loss is 0.82 mJ per pulse and

average power loss is 1.67 W.

The second stage diode was also measured with the same diagnostic setup as the

first stage diode loss measurements, and the loss was determined using the same

procedure as the first stage diode loss above. Figure 5.13 and Figure 5.14 displays the

voltage and current waveforms, along with the power loss waveform for the second stage

diode. The average power loss of 4.13 W was determined from 0.457 mJ integrated

energy loss per pulse. Since six diodes are connected in series, 690 mW per diode is

dissipated as heat. The temperature rise of these diodes is practically imperceptible.

Page 112: Research

97

Second Stage Diode Waveforms

-200

-150

-100

-50

0

50

100

150

200

10 11 12 13 14 15 16 17 18Time ( µs )

Dio

de V

olta

ge (

V )

-1

0

1

2

3

4

5

6

7

Dio

de C

urre

nt (

V )

Second stage diode voltage

Second stage diode current

Figure 5.13 Second stage diode voltage and current waveforms.

Second Stage Diode Power Waveform

-600

-400

-200

0

200

400

600

800

10 11 12 13 14 15 16 17Time ( µs )

Pow

er (

W )

Eloss=0.457 mJ

Figure 5.14 Second stage diode power loss waveform. Integrated energy loss is 0.457 mJ per pulse and

average power loss is 4.13 W.

Page 113: Research

98

5.2.4 Total Component Loss Calculation

In order to complete the total component loss calculation, the capacitor and

inductor loss must be determined. The capacitor loss is estimated by using the equivalent

series resistance, capacitor RMS current, and calculation of the I2R loss. The RMS value

of the first stage filter capacitor, C1, can be found by addition of the RMS current flowing

into C1 through D1, and current flowing out of C1 through L2. The RMS values are

calculated in a simple MATLAB routine from the measured current waveforms. The

current values calculated are

ID1 = 7.064 A rms and

IL2 = 3.55 A rms,

where ID1 is the RMS current into C1 and IL2 is the current out of C1. The MATLAB

routine for the RMS calculation is shown in Appendix D. The resultant C1 current equals

3.514 A rms. The current through C2 is 1.204 A rms using a similar procedure as with

the C1 current. Both capacitors have approximately 0.5 Ω series resistance.

The total capacitor power loss is given in Table 5.8, along with a summary of the

component losses presented in the previous sections. The inductor loss was determined

by neglecting stray system resistance loss, and assuming the inductors consume the

remaining loss. The following accounts for the total power loss of the inductor of each

stage, which includes core and I2R winding loss:

inductor total switch diode capacitorP P P P P= − − − , [5.5]

Page 114: Research

99

where Ptotal is the total power loss per stage, Pswitch is the IGBT loss, Pdiode is the diode

loss, and Pcapacitor is the capacitor loss as shown in Table 5.8 below.

Table 5.8 Average component power loss.

Total Loss per Stage

Switch Loss

Diode Loss

Capacitor Loss

Inductor Loss

First Stage 152.9 W 98.6 W 1.7 W 6.2 W 46.5 W

Second Stage 153.1 W 104.7 W 4.1 W 0.5 W 43.8 W

Total Average power ( W ) 306.0 W 203.3 W 5.8 W 6.7 W 90.3 W

5.3 Optimization of the Prototype Converter

The component loss analysis provides the insight as to where further design can

achieve a higher efficiency of the converter. According to Table 5.8, the switches and

inductors are the leading component losses of the prototype converter. As such, it is

believed that a loss mitigating design of the switches and inductors can significantly

improve the efficiency. In particular, the first stage inductor, made up of four inductors

in series in the prototype, can be re-designed as a single more efficient inductor, which

can utilize nanocrystalline core material similar to the second stage inductor design and

likely have reduced winding loss [4-5]. For the switches, minimizing the stray

inductance in the first and second stage IGBT connections will reduce voltage overshoot

during turn-off and the subsequent power loss. However, a significant improvement to

the switching loss can be made through dedicated research towards optimizing the turn-

off time. For instance, a thorough semiconductor trade study can reveal alternative

Page 115: Research

100

IGBTs with faster turn-off times for the first stage and second stage switch. For example,

the latest generation of fast-stop and light punch through IGBTs show evidence of faster

turn-off time than the non-punch through IGBTs in the prototype [6-7]. Also, faster

switching speed may be possible with experimentation with device configurations, such

as increasing the number of parallel IGBTs in the first stage or possibly even utilizing a

series and parallel matrix of high speed MOSFET devices. Additionally, development of

the gate drive circuitry, such as ultra-low impedance gate circuit can expedite gate charge

removal during the turn-off process. These efforts of lowering the switching loss and

optimizing the inductors have the potential for a future prototype converter design with

an efficiency as high as 90%. For this efficiency, the total loss of a 1 kW output

converter must be reduced to 111 W. The switching loss must be reduced by

approximately 75% and the inductor loss must be reduced by approximately 50% of the

presented prototype converter’s switching and inductor loss, respectively.

Page 116: Research

101

REFERENCES FOR CHAPTER 5.0 [1] V. K. Khanna, IGBT Theory and Design, Piscataway, NJ: IEEE Press, 2003.

[2] Cree Inc. application note; CPWR-AN03 (www.cree.com)

[3] G. Spiazzi, “Performance evaluation of a Schottky SiC power diode in a boost PFC application,” IEEE Trans. Power Electronics, vol. 18, issue 6, pp. 1249-1253, Nov. 2003.

[4] M. Ferch, “Light transformers for kilowatt SMPS based on nanocrystalline soft magnetic cores,” in Proc. of Seventh Int. Conf. on Power Electronics and Variable Speed Drives, 1998, pp. 411-415.

[5] W.A. Rease, “Design technology of high-voltage multi-megawatt polyphase resonant converter modulators” in Proc. of 29th Annu. Conf. IEEE Industrial Electronics Society, 2003, pp. 96-101.

[6] H, Husken, “FieldStop IGBT with MOS-like (tailless) turn-off,” in 2003 Proc. ISPSD, pp. 338-340.IGBT with MOS-like (tailless) turn-off,” in 2003 Proc. ISPSD, pp. 338-340.

[7] J. Yamada, “Low Turn-off Switching Energy 1200V IGBT Module,” in 2002 Proc. Industry Applications Conf., pp. 2165-2169.

Page 117: Research

102

CHAPTER 6.0 CASCADE BOOST CONVERTER SCALING TO HIGHER POWER LEVELS

The prototype cascade boost converter has provided a basis for a cascade boost

converter scaled to the higher voltage and power output levels of 60 kV and 300 kW, as

listed in Table 6.1, for power conditioning applications such as gyrotron microwave

tubes. As discussed in Chapter 1, gyrotron tubes require a high voltage source for

support of the high current electron gun in the depressed collector configuration. To

achieve this power conditioning output, the cascade boost circuit equations of Chapter 2,

and performance of the prototype converter have been integrated, and are used to derive

the higher power design. The DC input has also been selected as 1 kV for the converter

design, which requires a total converter step-up gain of 60 instead of 50, as in the

prototype converter. Multiple converters must be paralleled to achieve the constant

current, megawatt gyrotron tube power requirements of 2.4 MW provided by CPI in

Chapter 1. Thus, eight 300 kW converters modules must be paralleled to deliver 40 A, at

60 kV, to produce 2.4 MW of power. The following sections provide a near-term, 5-year,

and a 10-year volume and mass estimate of the cascade boost converter module, which

indicate a potential for a very compact and lightweight converter design.

Table 6.1 Cascade boost converter specifications for a 60 kV, 300 kW module.

Voltage Input 1000 V Voltage Output 60 kV Power Output 300 kW

Page 118: Research

103

6.1 Near-Term Technology Based Scaling

The design equations of Chapter 2 have been integrated into the design of the 60

kV, 300 kW cascade boost converter. The parameters calculated for an ideal 300 kW

cascade boost converter are shown in Table 6.2. The parameters of this system differ

from the 5 kV prototype system in correlation to the higher voltage step-up ratio of 60

and lower impedance load to deliver a 5A output. The load has been simplified as a

constant resistive load for a constant current electron beam, although in practice the load

can be modeled as a complex and time-varying impedance. These parameters are also

ideal parameters for a lossless converter, and have been calculated using synchronous

switching waveforms for both stages.

From the ideal converter design, the switching parameters must be modified for

the losses in an actual system. If the estimated efficiency of a 300 kW converter is

established to be 90%, which is believed to be a realistically achievable value based

primarily on mitigated switching losses as discussed in section 5.3, the total power loss

will be 33.3 kW. The parameters corresponding to a system with a 90% efficiency are

shown in Table 6.3. Thus, the parameters have been calculated to provide an average

input current of 333.3 A for an average input power of 333.3 kW. It has been assumed

that each stage contributes half of the power loss, similar to the results of the 5 kV

prototype cascade boost converter. For total system efficiency of 90%, each stage of the

300 kW converter must have an efficiency of approximately 95%. Therefore, the second

stage average input current has been adjusted to be 63.3 A for an average second stage

Page 119: Research

104

input power of 316.5 kW. The current waveforms through L1 and L2 indicate the peak

switch turn-off current of each stage, which is utilized for scaling the parallel diode and

switch assemblies and inductor core sizes. Also, due to the increasing duty ratio, the

capacitances C1 and C2 increase slightly; the adjusted capacitance values have been

determined to reflect the realistic capacitance values of a non-ideal converter.

Table 6.2 Ideal parameters calculated for a 60 kV, 300 kW cascade boost converter.

Switching frequency 10 kHz (both stages) Ideal duty ratio 0.6 (both stages) Load impedance 12 kΩ Average output current 5 A L1 74.9 µH L2 1.64 mH C1 110 µF C2 788 nF

Table 6.3 Switching parameters adjusted, corresponding to 90% efficiency.

Efficiency 90 % Average input power 333.3 kW Duty ratio D1 0.632 Duty ratio D2 0.617 L1 peak current 844 A L2 peak current 188 A C1 125 µF C2 1 µF

The following presents the methodology for the converter volume and mass

estimation using the values found in Table 6.3. The active switching components of the

first and second stages were designed as assemblies of series and parallel IGBTs similar

to the assemblies implemented in the 5 kV cascade boost converter prototype. The

primary switching device is based on the Powerex CM300DY-24NF IGBT module,

Page 120: Research

105

which has a rating of 1200 V and 300 A with two IGBTs integrated into a single IGBT

module package. This IGBT module is the latest generation of IGBT devices and

represents modern, high performance switching devices produced with silicon. This

IGBT uses trench-gate technology for reduced on-state voltage and the Light Punch

Through doping structure (discussed at the end of section 3.1.2) for fast charge carrier

recombination at turn-off, and is optimized for high switching frequency and low turn-off

loss. The device was selected based on published low turn-off energy loss figures and

high switching frequency ability [1]. The reported turn-off energy per pulse is 8.4 mJ

(0.028 mJ/A turn-off loss) with a bus voltage 600 V and switching 300 A [2]. Series and

parallel arrangement of these modules for S1 and S2 have low loss, where the turn-off

energy loss is considered the primary loss. The necessary number of devices in series

and parallel for S1 and S2 is shown in Table 6.4. The voltage and current per device

during turn-off and the average expected assembly turn-off loss, based on reported turn-

off energy loss per device, is shown in Table 6.5. Utilizing the reported energy loss

relation of 0.028 mJ/A and switching 281.3 A (shown in Table 6.5), the energy loss is

approximately 7.9 mJ per pulse. However, scaling the loss in relation to the higher

voltage of 833.3 V (instead of 600 V), scales it up by 38% to 11 mJ per pulse. The

switching frequency for the calculation of the turn-off loss is 10 kHz, which gives

approximately 110 W per IGBT.

Page 121: Research

106

Table 6.4 Number of IGBT modules in series and parallel for assembly of S1 and S2.

No. of series IGBTs

No. of parallel IGBTs

Total modules (2 IGBTs per module)

S1 6 3 9 S2 73 1 37

Table 6.5 Expected voltage and current per IGBT, and total average turn-off loss.

Voltage per IGBT

Current per IGBT

Average turn-off loss

S1 833.3 V 281.3 A 1980 W S2 810.8 V 188 A 5402 W

The cooling of the assemblies for the near-term is estimated based on COTS

liquid cooled heat sinks. On-board cooling systems of mobile platforms, such as aircraft

oil cooling, integrated with these liquid heat sinks will minimize the size and eliminate a

dedicated heat exchanger. The heat sink design is based on a flowed water component

from Aavid Thermalloy with 0.002 ºC/W thermal resistance rating for a 0.6”x7”x24”

model (part #: 416301U). The thermal resistance ratings of IGBT module cases and heat

sinks were used to maintain IGBT junction temperature below 125 ºC with the expected

power loss and liquid coolant temperature of 25 ºC. The junction temperature is found as

[6.1]

( )j d th lT P R T= + , [6.1]

where Pd is the average power loss, Rth is the total thermal resistance from IGBT chip

junction to the liquid coolant, and Tl is the liquid coolant temperature [3]. The volume

and mass of switches S1 and S2 (including heat sinks and IGBT assemblies) is included in

Page 122: Research

107

the component data for the first and second stage, which is shown in Table 6.6 and Table

6.7, respectively.

The inductor scaling method uses the second stage inductor dimensions of the 5

kV converter prototype and scales them up for the first and second stage inductors in the

60 kV, 300 kW converter design. This method provides an approximate estimate for the

projected inductor volume and mass. First, the peak flux density, Bm, of the inductor

cores is assumed to be conserved from the prototype as the inductor currents are scaled

up to 844 A and 188 A from the lesser prototype current levels for L1 and L2,

respectively. The peak flux density must be conserved since the inductor core must

operate below the magnetic saturation level of the nanocrystalline core material. Since

Bm is inversely proportional to the core area and directly proportional to the magnetic

flux, the relation becomes [6.2]:

mm

cB

= . [6.2]

where, Ac is the core cross section area and Φm is the maximum flux that occurs at peak

inductor current. If the inductor parameters are constant, then Bm is proportional to the

projected increased inductor current, by the following equation [6.3]:

4

g

0.4 102

l

DCm

c

IN IBm

A

π −∆⎛ ⎞+ ×⎜ ⎟Φ ⎝ ⎠= = [6.3]

Page 123: Research

108

where lg is the air gap length, N is the number of inductor winding turns, IDC is the DC

current level, and ∆I is the peak current level [4]. However, Bm can remain constant,

despite increases in the current level, by manipulating N and lg from equation [6.3] and

the core area Ac from equation [6.2]. If the inductor core area, Ac, is scaled by a factor

limited to approximately one-half of the scaling factor for peak inductor current value,

which requires a new winding density and longer air gap distance for increased core

reluctance, then the approximated scaling relation can be shown in equation [6.4]:

2

IA

SS = , [6.4]

where SA is the inductor core area scaling factor, and SI is the scaling factor for the peak

inductor currents from the prototype converter. The total volume and weight of the

inductor can then be scaled as the cube of the change in the length of one side of the

core’s cross-section area, based upon general inductor and transformer scaling trends [4].

If the core cross section area has an area scaling factor of SA as shown above, then the

length of one side of the core’s cross section is approximately proportional to the square-

root of SA. Thus, the inductor volume and mass scaling factor, SL, is equal to the

following equation [6.5]:

( )3L AS S= , [6.5]

Page 124: Research

109

where AS is the scaling factor of one of the sides of the core’s cross-section. The

dimension scaling factor SL allows for the inductor volume and mass scaling based on the

increase of inductor current in the 300 kW converter scaled up from the inductor current

of the 5 kV prototype converter. This method creates a scaling factor by conserving the

magnetic flux density, Bm in the first and second stage inductor cores, and by using the

second stage inductor in the 5 kV prototype as the baseline. The scaled inductor volume

and mass of L1 and L2 is shown in Table 6.6 and Table 6.7, respectively.

The diodes of the 60 kV converter have been scaled based on the diodes of the 5

kV converter prototype. The two parallel diodes of D1 in the prototype effectively form a

single diode module that is used as a model for a 500 V, 50 A rated diode module, and

the six series diodes of D2 in the prototype likewise form a 5000 V, 10 A diode module.

These diode modules, arranged in parallel and series, create the diodes for the 60 kV

converter. The volume and mass of D1 and D2 in the 60 kV converter are displayed in

Table 6.6 and Table 6.7, respectively.

Finally, the capacitors for near-term cascade boost converter were estimated

based upon the existing energy densities (J/cc) for high voltage capacitors. The

capacitors used in the 5 kV converter prototype did not provide the baseline for modern

technology, instead the energy density value of 0.47 J/cc and mass density value of 1.2

g/cc, of modern COTS capacitors were utilized for determining the mass and volume of

the capacitors in the converter. The capacitor data was taken from General Atomics

capacitor datasheets. The overall volume and mass was estimated based on these

Page 125: Research

110

volumes along with the amount of energy stored in C1 and C2. The energy stored in the

capacitors is found by [6.6]

212

storedE CV= (Joules), [6.6]

where V is the voltage across the capacitors. Using the capacitance value shown in Table

6.3, the energy stored in C1 (125 µF) is 1,562.5 J and the energy stored in C2 (1 µF) is

1,800 J. The volume and mass of C1 and C2 is shown in Table 6.6 and Table 6.7,

respectively.

Table 6.6 First stage components volume and mass for the near-term 60 kV, 300 kW converter.

in3 lbs S1 421.0 41.0 L1 565.2 46.0 D1 233.8 13.1 C1 202.9 8.8 Total 1,422.8 108.9

Table 6.7 Second stage components volume and mass for the near-term 60 kV, 300 kW converter.

in3 lbs S2 1,538.0 150.0 L2 3,348.0 96.0 D2 149.6 6.2 C2 233.7 10.1 Total 5,277.2 262.3

The component volume and mass data are displayed by the component category

in Table 6.8. The estimated mass of the near-term cascade boost converter shows that the

Page 126: Research

111

switches in the system consume about 50% of the mass, and the inductors consumes

about 38% of the converter mass.

Table 6.8 Estimated volume and mass of components of the near-term 60 kV, 300 kW converter.

in3 lbs Switches 1,959.0 191.0 Inductors 3,913.2 142.0 Diodes 383.3 19.3 Capacitors 436.6 18.9 Total 6,692.1 371.2

The total component volume and mass values in Table 6.9 do not describe a

completely packaged and operational 300 kW cascade boost converter. The total

component volume and mass values have been adjusted by a factor of 20% to estimate a

complete converter, accounting for the additional volume and mass of the converter

enclosure, shielding, support structure, control system electronics, insulation, and

possibly shock and vibration damping. Table 6.9 shows the estimated power density of

the near-term 60 kV, 300 kW converter using existing technology is approximately 1.5

lbs/kW.

Table 6.9 Total near-term projected component volume and mass and complete system volume and mass of a 300 kW converter. The 20% scaling factor estimates additional volume and mass for a completely

packaged and operational converter.

in3 lbs lbs/kW Total Component 6,692.1 371.2 Complete System

(20% factor) 8,030.5 445.4 1.5

Page 127: Research

112

6.2 Five & Ten-year Future Technology Scaling

Future technologies such as Silicon Carbide (SiC) semiconductor devices offer

the potential for great power density improvement of power conditioning systems.

Increased switching frequency and higher voltage handing per device are the leading

factors for the following projected reduction in volume and mass of the 300 kW cascade

boost converter module. The 1200 V SiC Schottky diode, implemented in the 5 kV

cascade boost converter prototype, already has shown advantages such as fast switching

speed and zero reverse recovery current [5-6]. Research has demonstrated SiC devices

will have a substantial impact on switch-mode converters and high voltage power

conditioning applications [7-9]. A switch-mode converter using 200 V SiC devices

(JFETs and diodes) has demonstrated reduced passive component size and higher power

density, which is largely due to approximately 6.7 times higher switching frequency than

a comparable converter with Si devices [10].

Properties of SiC that make the material so attractive for high voltage

semiconductor devices, compared to Si, include higher electric field breakdown, higher

thermal conductivity, larger saturation velocity, and higher operating temperature. The

SiC devices are known as wide band gap semiconductor devices—the band gap energy of

SiC is more than twice as large as Si. Table 6.10 lists the specific qualities of SiC in

comparison to Si. The band gap energy level affects the semiconductor material

operating temperature range and electric field breakdown strength. Two polytypes of SiC

that are used as semiconductor material: hexagonal 4H-SiC and 6H-SiC. The 4H-SiC

Page 128: Research

113

however is more commonly employed than 6H-SiC, which includes a higher band gap

and higher electron mobility.

Table 6.10 Properties of SiC that will enhance power semiconductor devices.

Si 4H-SiC Band Gap Energy 1.12 ev 2.2-2.9 ev Electric Field Breakdown 300 kV/cm 4000 kV/cm Thermal Conductivity ~ 1.5 W/cm-˚C ~ 5 W/cm-˚C Saturation Velocity ~ 1x107 cm/s ~ 2x107 cm/s Max Junction Temp. 150˚C 300˚C

The first generation of high voltage SiC devices used may not be IGBTs or

MOSFETs, instead SiC BJT power switching devices may enter the market first [11-14].

Apparently, gate oxide issues and low channel mobility coupled with material science

issues are the major challenges for high power SiC MOS controlled devices. These

challenges result in SiC gate structures which have reliability issues and high on-state

resistance in the inversion layer. Devices that do not use MOS control, such as the BJT,

eliminate such problems. High voltage, SiC BJT devices have recently been

demonstrated [15]. Also, BJTs and other minority carrier devices can theoretically

exhibit approximately 100 times faster switching speed than equivalent Si devices [16].

This is due to SiC’s 10 times higher electric field strength, which allows voltage blocking

layers to be made 0.1 times as thick. The ten-fold reduction in thickness of the voltage

blocking layers allows for a proportional reduction in the diffusion length of injected

minority carriers, and the shorter diffusion length reduces the carrier lifetime to 100 times

less. As discussed in Chapter 3, such reduction in carrier lifetime impacts the effect of

the current tail at turn-off, to be nearly eliminated. The SiC devices also offers thinner

Page 129: Research

114

base regions than Si, which allows SiC BJTs to operate with higher current gain (beta) for

reduced base drive current to control the device. The future SiC BJT, with fast switching,

high beta, and avoidance of problematic MOS control issues, becomes a strong candidate

for the first generation of commercial high voltage and high power SiC switching.

However, the SiC BJT may also share the same trend as the Si BJT ancestor. Ultimately,

with technical solutions of the aforementioned SiC material science challenges with SiC,

voltage controlled MOS SiC devices will be the next standard for solid state power

devices.

Reevaluation of the cascade boost converter using future SiC technology,

projected for the time range of 5 and 10 years, provides a technological basis for an

estimated size and weight reduction of the 300 kW cascade boost converter. The voltage

ratings and switching frequencies for these devices in 5 and 10 years have been

estimated. However, estimating the trend of SiC devices is highly unpredictable, as SiC

represents a new paradigm in semiconductors that is currently very young, and can only

represent a hypothesized technology projection. Therefore these estimates are based on

demonstrations, simulations, and projections found in published literature of SiC research

and development efforts [7] [10] [17-19]. A summary of the projected key parameter

estimates are shown in Table 6.11. The following 5-year trend estimate is that SiC

devices, implemented in the cascade boost converter, will achieve voltage blocking

ratings of 5 kV, and the switching speed will be increased to 70 kHz. The 10-year trend

estimate is that 8 kV ratings will exist per SiC device, and the switching frequency will

be increased to 100 kHz.

Page 130: Research

115

The faster switching ability of the SiC devices allows the converter to be operated

at the higher switching frequencies. However, the switching frequencies for the two time

frames are assumed to be limited by average switching power loss, which has been

assumed to be equal to the near-term average switching loss displayed in Table 6.5.

Thus, the average power loss in the semiconductor devices for all three time frames will

be assumed constant. This increase in switching frequency will reduce the volume and

mass of the inductors, which scale inversely proportional to the switching frequency. For

instance, the first stage inductor volume and mass has been scaled for the 5-year trend to

be one-seventh, or 14.3%, of the first stage inductor in the near-term because of the seven

times switching frequency increase. The ideal inductor derivation in Chapter 2 indicates

the linear relationship with the switching frequency.

Likewise, the capacitance value required for the first and second stage filter

capacitors will reduce linearly with the increase in frequency according the derivations in

Chapter 2. Subsequently, the energy storage requirement reduces linearly with the

capacitance in the 5 and 10-year time frames according to equation [6.6]. Also, energy

densities of 1 J/cc and 5 J/cc are anticipated to be available in the 5 and 10-year terms,

respectively, based upon reports in published literature [20-21]. The reduced energy

storage requirement, compounded with future high energy density dielectric technology,

result in dramatic size reduction of the capacitors. The first and second stage output

capacitors of the cascaded boost converter reduce to the resultant volume and mass of

approximately 6.7% of the near term, for the 5-year trend, and 1% of the near term, for

the 10-year trend. This advance in capacitor technology and the increase in switching

Page 131: Research

116

frequency reduce the first and second capacitors down to the size of circuit board

mountable components.

The higher voltage capability of SiC devices will reduce the necessary number of

devices in series for each switch stack assembly. The first and second stage switch

volume and mass are linearly scaled based on the number of devices, in the 5 and 10-year

assemblies, relative to the number of devices in the near-term switch assemblies. The

number of devices to be used in series and parallel for the first and second stage switch

assemblies, for the two time frames, is shown in Table 6.12.

Table 6.11 Summary of the key parameter estimates for SiC devices in 5 and 10 years.

5-year 10-year Voltage rating per device 5 kV 8 kV Converter switching freq. 70 kHz 100 kHz

Table 6.12 Number of devices to be placed in series and parallel for the 5 and 10-year time frames for the 300 kW converter’s first and second stage switches.

5-year (5 kV rated devices) 10-year (8 kV rated devices) Series Parallel Series Parallel

S1 2 3 1 3 S2 20 1 10 1 26 devices 13 devices

As discussed above, the average switching loss in the 5 and 10-year estimates will

remain equal to the near-term switching loss although fewer devices will be

implemented. Thus, the average power loss, on an individual device basis, will increase.

However, an estimation of future advanced cooling technology will allow for a favorable

scaling of the thermal management hardware, based on the literature available on

advanced cooling methods. The near-term switching assemblies have been designed with

Page 132: Research

117

conventional water cooled, bulk-aluminum heat sinks, with lower heat flux than

advanced technologies, which have been primarily motivated by the thermal management

needs of high performance server CPUs, VLSI devices, and high-density power

electronics. High heat flux removal from micro-channel systems for local heat extraction

has shown the ability to remove to up to 500 W/cm2 in local hotspots [22]. Future heat

flux may reach 1000 W/cm2 with the optimization of multi-phase fluid flow within

micro-channels for a number of future systems [23]. Furthermore, the near-term estimate

uses IGBT modules with standardized industrial packaging. Integrated, on-chip cooling

has been pursued for densely packed, high performance IC’s for highly efficient heat

extraction [22]. Customized packaging with integration of the cooling system within the

device package can eliminate the contact resistance between the IGBT base plate and the

heat sink. Finally, the typical maximum junction temperature of SiC is approximately

double the maximum junction temperature of Si, which reduces the required thermal

resistance of the cooling system by a factor of 2.

The approximated scaling, of the 300 kW converter’s cooling system volume and

mass, is assumed to be linear in relation to the number of switches in the 5 and 10-year

time frames. Therefore, the total switch assembly volume and mass will be directly

related to the reduction of the number of switches and diodes. For instance, the first stage

switch design in the 5-year trend will have a total mass of 13.7 lbs, or one-third of the

mass of the same assembly in the near-term, since there will be one-third the number of

devices in the assembly as in the near-term.

Page 133: Research

118

Table 6.13 and Table 6.14 show the first and second stage component volume and

mass for the 5-year and 10-year technology trend estimates. Table 6.15 shows a

breakdown per component of the cascade boost converter in the 5 and 10-year trends.

Table 6.16 and Table 6.17 give the estimates for the final system for the 5 and 10-year

trends including a 20% factor for additional volume and mass for a fully packaged and

operational converter. As with the near-term converter estimate, the 20 % accounts for

the additional hardware that may include the converter enclosure, shielding, support

structure, control system electronics, insulation, and possibly shock and vibration

damping.

Page 134: Research

119

Table 6.13 First stage component projected volume and mass for the 5 & 10-year trend in a 60 kV, 300 kW converter module.

5-year 10-year in3 lbs in3 lbs

S1 140.3 13.7 70.2 6.8 L1 80.7 6.6 56.5 4.6 D1 46.8 2.6 23.4 1.3 C1 13.6 0.6 1.9 0.1

Total 281.4 23.4 152.0 12.8

Table 6.14 Second stage component projected volume and mass for the 5 & 10-year trend in a 60 kV, 300

kW converter module.

5-year 10-year in3 lbs in3 lbs S2 421.4 41.1 209.7 20.4 L2 478.3 13.7 334.8 9.6 D2 52.5 2.2 26.2 1.1 C2 15.7 0.7 2.2 0.1

Total 967.8 57.7 573.9 31.1

Table 6.15 Estimated total volume and mass of each component category in a 60 kV, 300 kW converter

module.

5-year 10-year in3 lbs in3 lbs

Switches 561.7 54.8 280.9 27.4 Inductors 559.0 20.3 391.3 14.2 Diodes 99.2 4.8 49.6 2.4 Capacitors 29.3 1.3 43.7 1.9

Total 1,249.3 81.1 725.9 44.1

Table 6.16 5-year total projected component and complete system volume and mass for a 60 kV, 300 kW

converter module. The 20% scaling factor estimates additional volume and mass for a completely packaged and operational converter.

Page 135: Research

120

5-year scaled in3 lbs Lbs/kW

Total Component 1,249.3 81.1 Packaged System

(20% factor) 1,499.1 97.3 0.32

Table 6.17 10-year total projected component and total system volume and mass for a 60 kV, 300 kW

converter module. The 20% scaling factor estimates additional volume and mass for a completely packaged and operational converter.

10-year scaled in3 lbs Lbs/kW Total Component 725.9 44.1 Packaged System

(20% factor) 871.1 53.0 0.18

The volume and mass trends for a 300 kW converter in the near-term and for the 5

and 10-year projections are displayed in Figure 6.1 and Figure 6.2 respectively. Figure

6.3 and Figure 6.4 display the significant mass reduction trend for a 1.2 MW and 2.4 MW

system, where four 300 kW converters are paralleled for the 1.2 MW output and eight are

paralleled for the 2.4 MW power output. The mega-watt class power converter shows an

attractive scaling profile, where the projected technology advancements in high voltage

semiconductor devices can greatly increase the power density of high power conditioning

systems. The scaling indicates the potential of multi-megawatt class conversion under

500 lbs. The compact and lightweight potential of the cascade boost converter design

enables mega-watt class power support for applications such as gyrotrons aboard mobile

platforms.

Page 136: Research

121

Volume Scaling of a 300 kW Converter

871

1,499

8,031

0

1,000

2,000

3,000

4,000

5,000

6,000

7,000

8,000

9,000

Near-term 5-year 10-year

Vol

ume

( In

3 )

Figure 6.1 300 kW converter volume scaling from estimated technology trends.

Mass Scaling of a 300 kW Converter

97

445

53

0

100

200

300

400

500

Near-term 5-year 10-year

Mas

s ( L

bs )

Figure 6.2 300 kW converter mass scaling from estimated technology trends.

Page 137: Research

122

Volume Scaling of a 1.2 & 2.4 MW Converter

64,244

6,9693,4845,997

32,122 11,993

0

20,000

40,000

60,000

80,000

100,000

120,000

Near-term 5-year 10-year

Vol

ume

( In

3 )

2.4 MW Vol1.2 MW Vol

Figure 6.3 1.2 and 2.4 MW converter volume scaling using parallel 300 kW converter modules.

Mass Scaling of a 1.2 & 2.4 MW Converter

389

1,781

212

3,563

424

779

0

400

800

1,200

1,600

2,000

2,400

2,800

3,200

3,600

4,000

Near-term 5-year 10-year

Mas

s ( L

bs )

1.2 MW Mass2.4 MW Mass

Figure 6.4 1.2 and 2.4 MW converter mass scaling using parallel 300 kW converter modules.

Page 138: Research

123

REFERENCE FOR CHAPTER 6.0 [1] E.R. Motto and J.F. Donlon, “The latest advances in industrial IGBT module

technology,” in Proc. Applied Power Electronics Conference and Exposition, 2004, pp. 235-240.

[2] J. Yamada, “Low Turn-off Switching Energy 1200V IGBT Module,” in Proc. Industry Applications Conf., 2002, pp. 2165-2169.

[3] N. Mohan, Power Electronics, 2nd Edition. New York: John Wiley & Sons, 1995.

[4] Wm. T. McLyman, Magnetic Core Selection for Transformers and Inductors, New York and Basel: Marcel Dekker, Inc., 1964.

[5] A. Elasser, “A comparative evaluation of new silicon carbide diodes and state-of-the-art silicon diodes for power electronic applications,” IEEE Trans. on Industry Applications, vol. ED-39, pp. 915-921, July-Aug. 2003.

[6] G. Spiazzi, “Performance evaluation of a Schottky SiC power diode in a boost PFC application,” IEEE Trans. on Power Electronics, vol. ED-18, pp. 1249-1253, Nov. 2003.

[7] J. A. Cooper “SiC Power-Switching Devices—the Second Electronics Revolution?,” Proceedings of the IEEE, vol. ED-90, pp. 956–968, June 2002.

[8] A. Elasser and T.P. Chow, “Silicon carbide benefits and advantages for power electronics circuits and systems,” Proceedings of the IEEE, vol. ED-90, pp. 969-986, June 2002.

[9] J. Wang and B.W. Williams, “Evaluation of high-voltage 4H-SiC switching devices,” IEEE Trans. on Electronic Devices, vol. ED-46, pp. 589-597, March 1999.

[10] Abou-Alfotouch, “1 MHz hard-switched silicon carbide DC/DC converter.” in Proc. Applied Power Electronics Conference and Exposition, 2003, pp. 132-138.

[11] A.K. Agarwa, ”Recent progress in SiC bipolar junction transistors,” in Proc. Power Semiconductor Devices and ICs, 2004, pp. 361-364.

[12] Yanbin Lou, “High voltage (>1kV) and high current gain (32) 4H-SiC power BJTs using Al-free ohmic contact to the base,” IEEE Electronic Device Letters, vol. ED-

Page 139: Research

124

24, pp. 695-697, Nov. 2003.

[13] Y. Lou, “Fabrication and characterization of high current gain (Beta =430) and high power (23a-500V) 4H-SiC Darlington bipolar transistors,” IEEE Trans. on Electronic Devices, vol. ED-51, pp. 2211-2216, Dec. 2004.

[14] R. Sei-Hyung, “1800 V NPN bipolar junction transistors in 4H-SiC,” IEEE

Electronic Device Letter, vol. ED-22, pp. 124-126, March 2001.

[15] J. Zhang, “Demonstration of first 9.2 kV 4H-SiC bipolar junction transistor,” IEEE Electronic Device Letters, vol ED-40, pp. 1381-1382, Oct. 2004.

[16] T.R. McNutt, ”Silicon carbide PiN and merged PiN Schottky power diode models implemented in the Saber circuit simulator,” IEEE Trans. on Power Electronics, vol ED-19, pp. 573-581, May 2004.

[17] A.Q. Huang, “The future of bipolar power transistors,” IEEE Trans. on Electronic Devices, vol ED-48, pp. 2535-2543, Nov. 2001.

[18] C.M. Johnson, “Comparison of silicon and silicon carbide semiconductors for a 10 kV switching application,” in Proc. Power Electronics Specialists Conf., 2004, pp. 572-578.

[19] P.G. Neudeck, “High-temperature electronics - a role for wide bandgap semiconductors?,” Proceedings of the IEEE, vol. ED-90, pp. 1065-1076, June 2002.

[20] F.W. McDougal, “High energy density pulsed power capacitors,” in Digest of Technical Papers of Pulsed Power Conf., 2003, vol. 1, pp. 513-517, June 2003.

[21] M.V. Fazio, “Ultracompact Pulsed Power,” Proceedings of the IEEE, vol. 92, pp. 1197-1204, July 2004.

[22] G. Upadhya, “Closed-loop cooling technologies for microprocessors,” in Tech Digest IEEE Int. Electron Devices Meeting, 2003, pp. 32.4.1 - 32.4.4.

[23] D. Faulkner, “Practical design of a 1000 W/cm2 cooling system,” in Proc. 19th Annu. IEEE Semiconductor Thermal Measurement and Management Symposium, 2003, pp. 223-230.

Page 140: Research

125

CHAPTER 7.0 CONCLUSION

This work has provided the design of the cascade boost converter topology, which

has been experimentally demonstrated as a viable converter for high step up ratio, high

voltage, DC-DC conversion through a prototype cascade boost converter. The equations

derived, successfully enabled the design, fabrication, and testing of the prototype cascade

boost converter by producing a 5 kV DC, 1 kW output from a constant 100 V input at an

efficiency of 76.6%. Furthermore, with the fruition of anticipated high voltage silicon

carbide power devices and high energy density capacitor dielectrics, the power density of

high voltage power conditioning systems will drastically increase. The cascade boost

converter topology has been shown to have very attractive scaling trends in response to

these technological advances for the 5 and 10-year projections.

The cascade boost converter prototype was designed based on the theoretical

converter work as described in Chapter 2. After the circuit analysis of the prototype, the

switches and inductors appear to be the leading sources of power loss in the converter.

Chapter 3 presented a discussion on IGBT switching considerations, where the

application of IGBTs must take into consideration switching loss, with respect to the

voltage rating of the IGBTs. Six series, 1200 V rated IGBTs implemented for the second

stage switch, enabled a 76.6% total prototype efficiency, and was compared to two series

3.3 kV rated IGBTs, as the second stage switch, which resulted in a lower prototype

efficiency of 37%. As discussed in Chapter 5, further research on semiconductor devices,

with optimization of series and parallel arrangement, will likely extend high voltage

Page 141: Research

126

capability and mitigate switching losses in future designs of the cascade boost converter.

Also, dedicated inductor optimization may reduce copper winding and magnetic core

hysteresis loss of the first and second stage inductors. An efficiency reaching 90% may

likely be realized with dedicated component optimization and circuit fabrication, which

has been based on testing of the cascade boost prototype converter.

Using the assumption that 90% efficiency is possible, system scaling in Chapter 6

for a 60 kV and 300 kW converter for the near-term, 5-year, and 10-year time frames

have been estimated. The scaling uses the basic series and parallel switching concepts

implemented by the prototype cascade boost converter. The COTS IGBT module used

for the first and second stage switch assemblies, in the 60 kV converter design, is a 1200

V, 300 A device that exhibits low turn-off loss, and has been optimized to mitigate the

current tail by using Light Punch Through technology. The diodes have been estimated

by scaling the necessary number of diodes in series and parallel to accomplish the diode

parameters for the higher voltage and power level. The inductors were scaled based on

the core cross sectional area required to prevent magnetic flux saturation in the cores due

to the peak inductor current, scaled up from the prototype converter. Chapter 6 discusses

estimated volume and mass of a converter scaled to 60 kV and 300 kW in the near-term,

5-year, and 10-year time frames. The 10-year converter estimation shows a potential for

a 300 kW system volume under 1,000 in3 and approximately 53 lbs. Parallel 300 kW

modules can be utilized to fabricate multi-megawatt class converters. For instance, four

converter modules assembled in parallel for a 1.2 MW system, scales to approximately

Page 142: Research

127

3,500 in3 and 212 lbs, and a 2.4 MW system will proportionally have a volume of 7,000

in3 and a mass of 424 lbs, remarkable by any system standards.

The cascade boost converter shows that reduction by 88% in weight may be

achievable in a 10-year time frame, which allows for high power conditioning systems to

be integrated onto platforms that are impractical today, such as mobile platforms. High

energy lasers, weapon systems, and pulsed power systems for the military can be

operated on land vehicles, aircrafts, and possibly man-portable systems given this

technology trend. Such power density advances will allow power conditioning systems

to satisfy volume and mass requirements for such military applications.

Page 143: Research

128

APPENDIX A

Drive Circuit Schematics of the Second Stage IGBT Stack

Below, Figure A.1 through Figure A.3, respectively show circuit schematics of

the trigger module, power oscillator, and gate drive circuit in the IGBT stack of six

individual, 1200V rated, IGBTs that is used in the second stage of the cascade boost

converter prototype. The IGBT trigger module, in Figure A.1, displays the six fiber optic

transmitters (HFBR1521), synchronized with the 50 Ω trigger TTL input signal, that send

optical signals to each IGBT gate driver board. The gate driver boards are powered by

the power oscillator shown in Figure A.2, where the secondary winding of T1 couples to

a magnetic core (CST206-1A) on each IGBT driver board, shown in Figure A.3, using a

single loop of high voltage insulated wire that is strung through each driver board. On

each gate driver board, the fiber optic receiver (HFBR2521) is coupled to the main gate

driver IC (MIC4451) that provides +15 V for turn-on and 0 V for turn-off with the

capability of 12 A peak drive current. Also, a 4.7 Ω external gate resistor is in series with

the IGBT (IRGP30B120KD-E) gate input, and each IGBT is overvoltage protected by a

series of six transient voltage suppressor diodes (D10-D15) that provide a clamping

voltage of 1200 V for the IGBT collector to the emitter voltage.

Page 144: Research

129

Figure A.1 IGBT trigger module.

Figure A.2 Power oscillator.

Page 145: Research

130

Figure A.3 IGBT gate driver circuit.

Page 146: Research

131

APPENDIX B

MATLAB function integratcsv.m. This routine integrates waveform data saved as an array in a comma separated

variable (.csv) file. The routine is specifically designed to open a file with two columns

of data. Column “B” is numerically integrated with respect to column “A”, where

column A may represent time values and column B represent waveform data. When

implemented in MATLAB, the result will be the integral of the data. Integratecsv.m is

shown below:

function Intgrl = integratecsv(csvfile) %integratecsv('directory path\filename.csv') %integratecsv numerically integrates a csv file (comma sep var) where %column B is integrated with respect to column A. M = csvread(csvfile); t = M(:,1); x = M(:,2); Y = trapz(t,x); Intgrl = Y(end);

Page 147: Research

132

APPENDIX C

Derivation of the IGBT conduction power loss

The conduction power loss of the IGBTs is found by multiplication of the on-state

current and voltage for a single period during stead state operation. The solution for the

power is integrated to give the energy loss during the on-state per period and multiplied

by the switching frequency, which provides the average conduction power loss. The

conduction current is given by [C.1]

( ) Linitial

VIcond t t IL

= + , [C.1]

where VL is the voltage across the inductor, L is the inductance, and Iinitial is zero if the

boost converter is in discontinuous mode conduction. The on-state voltage, vf, of the

IGBT is approximated as [C.2]

0.7 0.068 Lf

Vv tL

= + , [C.2]

which was determined from on-state forward conduction test shown in

Figure C.1. Multiplication of [C.1] and [C.2] results in the power equation: [C.3]

2

( ) 0.7 0.068L Lcond

V VP t t tL L

⎛ ⎞= + ⎜ ⎟⎝ ⎠

. [C.3]

Integration of Pcond, results in the conduction energy loss per period: [C.4]

( ) ( )2

2 3

0

1 1( ) 0.7 0.0682 3

DTL L

on condV VE P t DT DTL L

⎛ ⎞ ⎛ ⎞ ⎛ ⎞= = +⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎝ ⎠ ⎝ ⎠ ⎝ ⎠∫ , [C.4]

Page 148: Research

133

where the integration limits are from the beginning of the period to end of the switch on-

time, DT. The average of Pcond(t) equals the Eon multiplied be the switching frequency:

[C.5]

1cond sw on onP f E E

T= = . [C.5]

On-State Forw ard Voltage Data & Trendline

0

10

20

30

40

50

60

0 1 2 3 4 5 6Vf

Am

ps

Vf data

Poly. (Vfdata)

Vg = 12.5 V

Figure C.1 Experimental on-state forward conduction voltage and trend line of the International Rectifier

IRGP30B120KD-E 1200 V, 60 A rated, Non-Punch Through (NPT), IGBT; Vg = 12.5 V.

Page 149: Research

134

APPENDIX D

MATLAB function RMSCalc.m. This routine calculates the RMS value of data saved as an array in a comma

separated variable (.csv) file. The routine is specifically designed to open a file with two

columns of data, where column A may represent time values and column B represent

waveform data. The time duration of the data must be equal to one period of a periodic

waveform. The result of the routine is the calculated RMS value of the data.

RMSCalc.m is shown below:

function Irms = rmscalc(csvfile) %rmscalc('directory path\filename.csv') %rmscalc calculates the RMS value of a waveform in a CSV file, particularly from a Tektronics oscilloscope. %First and last data point in the CSV file define the period of a repetitive waveform. M = csvread(csvfile); t = M(:,1); x = M(:,2); Y = cumtrapz(t,x.^2); T = max(t)-t(1); Irms = sqrt(Y(end)/T);