the design of a micro-turbogenerator

144
The Design of a Micro-turbogenerator by Andrew Phillip Camacho Department of Mechanical Engineering and Materials Science Duke University Date: Approved: Jonathan Protz, Supervisor Devendra Garg Rhett George Thesis submitted in partial fulfillment of the requirements for the degree of Master of Science in the Department of Mechanical Engineering and Materials Science in the Graduate School of Duke University 2011

Upload: others

Post on 07-Nov-2021

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The Design of a Micro-turbogenerator

The Design of a Micro-turbogenerator

by

Andrew Phillip Camacho

Department of Mechanical Engineering and Materials ScienceDuke University

Date:

Approved:

Jonathan Protz, Supervisor

Devendra Garg

Rhett George

Thesis submitted in partial fulfillment of the requirements for the degree ofMaster of Science in the Department of Mechanical Engineering and Materials

Sciencein the Graduate School of Duke University

2011

Page 2: The Design of a Micro-turbogenerator

Abstract(micro-turbogenerator design)

The Design of a Micro-turbogenerator

by

Andrew Phillip Camacho

Department of Mechanical Engineering and Materials ScienceDuke University

Date:

Approved:

Jonathan Protz, Supervisor

Devendra Garg

Rhett George

An abstract of a thesis submitted in partial fulfillment of the requirements forthe degree of Master of Science in the Department of Mechanical Engineering and

Materials Sciencein the Graduate School of Duke University

2011

Page 3: The Design of a Micro-turbogenerator

Copyright c© 2011 by Andrew Phillip CamachoAll rights reserved except the rights granted by the

Creative Commons Attribution-Noncommercial Licence

Page 4: The Design of a Micro-turbogenerator

Abstract

The basic scaling laws that govern both turbomachinery and permanent magnet gen-

erator power density are presented. It is shown for turbomachinery, that the power

density scales indirectly proportional with the characteristic length of the system.

For permanent magnet generators, power density is shown to be scale independent

at a constant current density, but to scale favorably in actuality as a result of the

scaling laws of heat dissipation.

The challenges that have affected micro-turbogenerators in the past are presented.

Two of the most important challenges are the efficiency of micro-turbomachinery and

the power transfer capabilities of micro-generators.

The basic operating principles of turbomachinery are developed with emphasis on

the different mechanisms of energy transfer and how the ratio of these mechanisms in

a turbine design relates to efficiency. Loss models are developed to quantify entropy

creation from tip leakage, trailing edge mixing, and viscous boundary layers over

the surface of the blades. The total entropy creation is related to lost work and

turbine efficiency. An analysis is done to show turbine efficiency and power density

as a function of system parameters such as stage count, RPM, reaction, and size.

The practice of multi-staging is shown to not be as beneficial at small scales as

it is for large scales. Single stage reaction turbines display the best efficiency and

power density, but require much higher angular velocities. It is also shown that for

any configuration, there exists a peak power density as a result of competing effects

iv

Page 5: The Design of a Micro-turbogenerator

between the scaling laws and viscous losses at small sizes.

The operating principles of generators and power electronics are presented as

are the scaling laws for both permanent magnet generators and electro-magnetic

induction generators. This analysis shows that permanent magnet generators should

have higher power densities at small sizes. The basic concepts of permanent magnet

operation and magnetic circuits are explained, allowing the estimation of system

voltage as a function of design parameters. The relationship between generator

voltage, internal resistance, and load power is determined.

Models are presented for planar micro-generators to determine output voltage,

internal resistance, electrical losses, and electromagnetic losses as a function of geom-

etry and key design parameters. A 3 phase multi-layer permanent magnet generator

operating at 175,000 RPM with an outer diameter of 1 cm is then designed. The

device is shown to convert 10 W of input shaft power into DC electric power at an

efficiency of 64%. A second device is designed using improved geometries and system

parameters and operates at an efficiency of 93%.

Lastly, an ejector driven turbogenerator is designed, built, and tested. A thermo-

dynamic cycle for the system is presented in order to estimate system efficiency as

a function of design parameters. The turbo-generator was run at 27,360 RPM and

demonstrated a DC power output of 7.5 mW.

v

Page 6: The Design of a Micro-turbogenerator

Contents

Abstract iv

List of Tables ix

List of Figures x

Acknowledgements xv

1 Introduction 1

1.1 Background and Motivation . . . . . . . . . . . . . . . . . . . . . . . 1

1.1.1 Concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1.2 Challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

1.1.3 Review of Previous Work . . . . . . . . . . . . . . . . . . . . . 22

1.2 Research Objectives, Goals, and Design Overview . . . . . . . . . . . 31

2 Micro-turbine Design and Performance 33

2.1 Concepts of Turbomachinery . . . . . . . . . . . . . . . . . . . . . . . 33

2.1.1 Operating Concepts . . . . . . . . . . . . . . . . . . . . . . . . 33

2.1.2 Conservation Laws and Governing Equations . . . . . . . . . . 37

2.1.3 Turbine Reaction, Flow Type, and Stage Count . . . . . . . . 39

2.2 Micro-turbine Design . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

2.2.1 Turbine Efficiency and Loss Accounting . . . . . . . . . . . . . 42

2.2.2 Characterization of the Flow . . . . . . . . . . . . . . . . . . . 47

2.2.3 Boundary Layer Losses . . . . . . . . . . . . . . . . . . . . . . 48

vi

Page 7: The Design of a Micro-turbogenerator

2.2.4 Trailing Edge Losses . . . . . . . . . . . . . . . . . . . . . . . 51

2.2.5 Tip Clearance Losses . . . . . . . . . . . . . . . . . . . . . . . 54

2.2.6 Scaling Effects on the Efficiency of Gas Turbines . . . . . . . . 57

2.3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

3 Permanent Magnet Micro-generator Design and Performance 68

3.1 Concepts of Permanent Magnet Generators . . . . . . . . . . . . . . . 68

3.1.1 Operating Concepts, Scaling Laws, and Generator Selection . 68

3.1.2 Magnetic Circuits and Modeling . . . . . . . . . . . . . . . . . 76

3.1.3 Electric Circuits and Modeling . . . . . . . . . . . . . . . . . . 86

3.1.4 Magnetic Materials and Configuration . . . . . . . . . . . . . 89

3.2 Device Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

3.2.1 Device Geometry and Parameters . . . . . . . . . . . . . . . . 94

3.2.2 Flux and Induced Voltage . . . . . . . . . . . . . . . . . . . . 99

3.2.3 Generator Coil Resistance . . . . . . . . . . . . . . . . . . . . 102

3.2.4 Maximum Power Transfer Capabilities . . . . . . . . . . . . . 104

3.2.5 Loss Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . 105

3.2.6 Power Output, Efficiency, and Layer Count . . . . . . . . . . . 110

3.3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

4 Experimental Results of an Ejector Driven Micro-turbogenerator 113

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

4.2 Thermodynamic Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . 114

4.3 Experiment and Results . . . . . . . . . . . . . . . . . . . . . . . . . 116

4.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

5 Conclusion and Future Work 121

5.1 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . 121

vii

Page 8: The Design of a Micro-turbogenerator

5.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

Bibliography 125

viii

Page 9: The Design of a Micro-turbogenerator

List of Tables

3.1 Voltage results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

3.2 Maximum power dissipation . . . . . . . . . . . . . . . . . . . . . . . 105

3.3 Eddy current losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

3.4 Ohmic losses in the stator windings and power electronics . . . . . . . 109

3.5 Load power and efficiency as a function of layer count . . . . . . . . . 110

4.1 Efficiency approximations . . . . . . . . . . . . . . . . . . . . . . . . 117

4.2 Experimental results . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

ix

Page 10: The Design of a Micro-turbogenerator

List of Figures

1.1 A micro-turbine engine operating a Brayton cycle . . . . . . . . . . . 3

1.2 Scaling laws for heat dissipation in a conductor . . . . . . . . . . . . 6

1.3 Energy density of micro-turbogenerators relative to Li-ion batteries . 7

1.4 TS diagrams for a Brayton cycle . . . . . . . . . . . . . . . . . . . . . 8

1.5 Gas flow path for a micro-turbine . . . . . . . . . . . . . . . . . . . . 10

1.6 Velocity triangles of a turbine stage . . . . . . . . . . . . . . . . . . . 11

1.7 Thermal resistance structures . . . . . . . . . . . . . . . . . . . . . . 12

1.8 A magnetizing head . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

1.9 Leakage flux permeating incorrect sectors during magnetization . . . 14

1.10 Trapezoidal magnetization profile of an annular permanent magnet . 14

1.11 Discrete magnet pieces arranged peripherally around the generator rotor 15

1.12 A silicon frame for discrete peripherally arranged magnet pieces . . . 15

1.13 Coil winding patterns for 2D micro-generators . . . . . . . . . . . . . 17

1.14 Silicon laminations within the stator back iron to reduce eddy currents 18

1.15 Temperature effects on magnetization . . . . . . . . . . . . . . . . . . 18

1.16 A single-zone combustor schematic . . . . . . . . . . . . . . . . . . . 19

1.17 Schematic of a dual zone combustor . . . . . . . . . . . . . . . . . . . 21

1.18 A catalytic micro-combustor . . . . . . . . . . . . . . . . . . . . . . . 21

1.19 A 10 mm axial impulse turbine . . . . . . . . . . . . . . . . . . . . . 22

1.20 SEM photograph of the MEMS turbocharger . . . . . . . . . . . . . . 23

x

Page 11: The Design of a Micro-turbogenerator

1.21 Cross sectional view of the turbomachinery . . . . . . . . . . . . . . . 23

1.22 A bird’s eye SEM photograph of the MEMS turbine . . . . . . . . . . 23

1.23 A split SEM view showing the concentric turbine stages . . . . . . . . 24

1.24 A fully packaged micro-combustor . . . . . . . . . . . . . . . . . . . . 25

1.25 Schematic of a dual zone combustor . . . . . . . . . . . . . . . . . . . 25

1.26 A catalytic combustor schematic . . . . . . . . . . . . . . . . . . . . . 26

1.27 Schematic of a ball bearing supported turbo-generator . . . . . . . . 26

1.28 Original and optimized winding patterns . . . . . . . . . . . . . . . . 27

1.29 A comparison between SmCo and NdFeB generators as a function oftemperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

1.30 Power density and ejector efficiency as a function of the nozzle throatdiameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

1.31 Suction draft performance versus entrainment ratio . . . . . . . . . . 29

1.32 A schematic of a MEMS power plant utilizing a Rankine cycle . . . . 29

1.33 Cross sectional schematic of the fully integrated permanent magnetturbogenerator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

1.34 The original turbogenerator and power electronics . . . . . . . . . . . 31

2.1 A control volume for a general turbo-machine . . . . . . . . . . . . . 33

2.2 Velocity triangles for the exit flow from a rotor blade . . . . . . . . . 34

2.3 A fluid element in a centrifugal field . . . . . . . . . . . . . . . . . . . 36

2.4 Velocity triangles for an impulse turbine rotor blade row . . . . . . . 39

2.5 Velocity triangles for a reaction turbine rotor blade row . . . . . . . . 40

2.6 Multistage turbine arrangement . . . . . . . . . . . . . . . . . . . . . 41

2.7 A radial micro-turbine . . . . . . . . . . . . . . . . . . . . . . . . . . 41

2.8 Effectiveness of impulse turbines and a reaction turbine as a functionof the velocity ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

2.9 An irreversible flow process through a throttle . . . . . . . . . . . . . 46

xi

Page 12: The Design of a Micro-turbogenerator

2.10 Simplified velocity triangles for idealized impulse and reaction blades 48

2.11 Dissipation coefficients for various boundary layers . . . . . . . . . . . 50

2.12 Wake mixing behind the trailing edge of two turbine blades . . . . . . 51

2.13 A dump diffuser . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

2.14 Total pressure drop in a diffuser . . . . . . . . . . . . . . . . . . . . . 53

2.15 Thick trailing edges of a micro-turbine . . . . . . . . . . . . . . . . . 55

2.16 A schematic of tip leakage over a turbine blade . . . . . . . . . . . . 56

2.17 Turbine effectiveness for a single stage reaction turbine . . . . . . . . 58

2.18 Turbine effectiveness for a single stage impulse turbine . . . . . . . . 59

2.19 Turbine effectiveness for a two stage impulse turbine . . . . . . . . . 59

2.20 Turbine effectiveness for a three stage impulse turbine . . . . . . . . . 60

2.21 Turbine efficiency for a single stage reaction turbine . . . . . . . . . . 60

2.22 Turbine efficiency for a single stage impulse turbine . . . . . . . . . . 61

2.23 Turbine efficiency for a two stage impulse turbine . . . . . . . . . . . 61

2.24 Turbine efficiency for a three stage impulse turbine . . . . . . . . . . 62

2.25 Power density for a single stage reaction turbine . . . . . . . . . . . . 64

2.26 Power density for a single stage impulse turbine . . . . . . . . . . . . 64

2.27 Power density for a two stage impulse turbine . . . . . . . . . . . . . 65

2.28 Power density for a three stage impulse turbine . . . . . . . . . . . . 65

3.1 A current induced magnetic field . . . . . . . . . . . . . . . . . . . . 69

3.2 A permanent magnet used to create a magnetic field . . . . . . . . . 70

3.3 The induced voltage being used to create a current and produce work 70

3.4 A three phase power system . . . . . . . . . . . . . . . . . . . . . . . 71

3.5 A phasor diagram for a wye connected generator . . . . . . . . . . . . 72

3.6 A diode bridge rectifier . . . . . . . . . . . . . . . . . . . . . . . . . . 72

xii

Page 13: The Design of a Micro-turbogenerator

3.7 The original AC voltage forms and the rectified waveform . . . . . . . 73

3.8 Rectifier current pathways . . . . . . . . . . . . . . . . . . . . . . . . 73

3.9 Scale effects on basic magnet interactions . . . . . . . . . . . . . . . . 75

3.10 Scale effects on magnetic interactions with increased current density . 76

3.11 A B-H graph for a hard magnetic material . . . . . . . . . . . . . . . 77

3.12 A simple permanent magnet circuit . . . . . . . . . . . . . . . . . . . 77

3.13 A simple permanent magnet circuit with an air gap . . . . . . . . . . 78

3.14 The B-H operating conditions for a permanent magnet system . . . . 80

3.15 A toroidal ferromagnetic core wrapped within a current conducting wire 80

3.16 A close up of the B-H curve for a hard magnetic material . . . . . . . 81

3.17 A circuit model for a permanent magnet . . . . . . . . . . . . . . . . 83

3.18 A circuit model for a permanent magnet enclosure with an air gap . . 83

3.19 A circuit model for a planar permanent magnet generator . . . . . . . 84

3.20 A simplified circuit model for a planar permanent magnet generator . 84

3.21 A B-H curve for a permanent magnet material undergoing magnetization 85

3.22 A magnetizing head used to create permanent magnets . . . . . . . . 86

3.23 A single phase AC circuit . . . . . . . . . . . . . . . . . . . . . . . . 86

3.24 Load power as a function of efficiency . . . . . . . . . . . . . . . . . . 89

3.25 Demagnetization curves for various magnetic materials . . . . . . . . 90

3.26 A comparison between SmCo and NdFeB generators as a function oftemperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

3.27 An annular magnet and mounting adaptor . . . . . . . . . . . . . . . 92

3.28 Discrete magnet pieces arranged peripherally around the generator rotor 93

3.29 A B-H curve for a magnetic material . . . . . . . . . . . . . . . . . . 94

3.30 The arrangement of an uncoupled micro-turbogenerator . . . . . . . . 94

3.31 A permanent magnet generator with three coil layers . . . . . . . . . 96

xiii

Page 14: The Design of a Micro-turbogenerator

3.32 A representational circuit diagram for a generator with a variable layercount, n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

3.33 Effect of layer count on system parameters . . . . . . . . . . . . . . . 97

3.34 A single phase of the 4 pole-pair, 3 turns per pole stator . . . . . . . 98

3.35 Planar continuum layers for a planar PM generator . . . . . . . . . . 99

3.36 A magnetization profile for an annular permanent magnet . . . . . . 100

3.37 Back iron thickness schematic . . . . . . . . . . . . . . . . . . . . . . 101

3.38 Laminations embedded within a magnetic conducting stator . . . . . 106

3.39 Eddy currents within a radial conducting segment with and withoutlaminations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

3.40 A representational circuit for power balance . . . . . . . . . . . . . . 109

3.41 Energy density of designed micro-turbogenerators relative to Li-ionbatteries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

4.1 Ejector control volume . . . . . . . . . . . . . . . . . . . . . . . . . . 114

4.2 After-burning thermodynamic cycle . . . . . . . . . . . . . . . . . . . 115

4.3 The original micro-turbogenerator . . . . . . . . . . . . . . . . . . . . 117

4.4 3D printed turbo-generator . . . . . . . . . . . . . . . . . . . . . . . . 118

4.5 LEDs powered by the micro-turbine generator . . . . . . . . . . . . . 119

4.6 Turbine shaft power and load power as a function of rotor speed . . . 120

xiv

Page 15: The Design of a Micro-turbogenerator

Acknowledgements

I would first like to thank my advisor, Dr. Jonathan Protz, for all of his help and

guidance over these past few years. With the exception of my parents, no one has

ever done as much for me. I would also like to thank my parents and sisters for their

support over this time. Most of this work is a direct result of the work of others

before me at MIT and GaTech, and I wish to strongly acknowledge that and thank

all of them for contributing to the scientific literature on this subject. I want to

thank Will Gardner and Ivan Wang for not only greatly assisting me academically,

but for also being good friends throughout this period. I also wish to thank Yuxuan

Hu for her emotional support over the years. You’ve always been there for me, both

during the highs and the lows. And while I will forever deny it, everyone claims that

you motivated me to become a hard worker and a better person. Things did not work

out between us, but I wish you my best. I know you will have a happy, fulfilling,

and successful life. I also want to thank Gwen Gettliffe for being by my side during

much of this time and giving me a reason to keep on going. You are a wonderful

person in so many ways, and I wish things had worked out differently between us,

but I take comfort in knowing that at least we will continue to be good friends. A

strong degree of gratitude goes to Kathy Parrish for simplifying all of the graduate

school requirements, I would have been lost without you. I also want to thank all of

my professors over the years for being patient and allowing me to build a stronger

fundamental understanding of some very abstract concepts, in particular Dr. George

xv

Page 16: The Design of a Micro-turbogenerator

for his E.E. insights and Dr. Shaughnessy for his compressible flow insights. I also

want to thank Dr. Garg for allowing me to pursue my interest in robotics during this

time. Kip Coonley and Justin Miles were especially kind in allowing me to use their

electronic equipment, sensors, and SMT soldering equipment. I would not have been

able to do my experiments without that support, thank you. Lastly, I would like to

thank Logos Technologies, DARPA, and Duke University for financial support over

this period.

xvi

Page 17: The Design of a Micro-turbogenerator

1

Introduction

1.1 Background and Motivation

1.1.1 Concept

In the mid 1990’s, the idea of using micro-sized devices to produce electrical power

from combustible fuels was presented by Dr. Epstein et al. (1997). The intent was

to use micro-sized heat engines in lieu of chemical batteries or fuel cells in specific

applications due to certain advantages that micro-heat engines would posses. To

demonstrate this, a quick synopsis of the strengths and weaknesses of these different

power sources must be given.

Chemical batteries are typically associated with high power densities but low

energy densities. As an example, conventional lithium-ion batteries are capable of

high power densities in the 1000 W/kg range, but also typically exhibit low energy

densities in the 150 W-hr/kg range [Panasonic Corporation (2009)]. In contrast,

fuel cells typically demonstrate high energy density due to the large specific energy

of their fuels and a good conversion efficiency, but also low power densities. A

compilation of commercial fuel cells by Narayan and Valdez (2008) showed energy

densities as high as 805 W-hr/kg, but with an accompanying specific power of only

1

Page 18: The Design of a Micro-turbogenerator

2 W/kg. Other fuel cells were shown that demonstrated power densities as high as

18 W/kg, but this occurred at the expense of a much lower conversion efficiency as

demonstrated by a specific energy of 121 W-hr/kg, less than a lithium ion battery.

Thus fuel cells and batteries exist at opposite ends of a trade-off spectrum, with

batteries demonstrating low energy density but high power density, and fuel cells

demonstrating high energy density but low power density. In between these two

extremes are hydrocarbon combustion based heat engines. As an example, a 1000 W

Honda generator using 2 gallons of gasoline has a specific energy of 506 W-hr/kg and

a specific power of 40 W/kg [Honda Corporation (2001)]. Reasonably high energy

densities can be achieved despite a low conversion efficiency due to the large specific

energy of hydrocarbon fuels (12,500 W-hr/kg for gasoline).

As a result of these trade-offs, these device classes are each well suited to different

applications. In particular, in cases where reasonably high power and energy densities

must be obtained such as in automobiles and aircraft, heat engines are most often

used. This trend continues down to the size where relatively cheap and conventional

machining practices can be utilized. As an example, small sized piston and gas

turbine engines are used for long endurance RC hobby aircraft. Below such sizes

however, relatively cheap manufacturing processes with the necessary precision no

longer exist. For this reason, batteries and fuel cells are currently used exclusively

at these smaller sizes for applications that would be better suited for a heat engine.

This work discusses the design of a heat engine, a micro-turbogenerator, that

could be manufactured to fill this void. Its operation is analogous to a convention-

ally sized natural gas power plant, where a gas turbine utilizing a Brayton cycle is

driven by the combustion of fuel, and the resulting shaft power is outputted to an

electric generator to create electricity. A figure of the engine core without the elec-

trical components can be seen in figure 1.1. At the macro-size level, gas turbines are

excellent devices for converting heat from combustion into shaft power due to their

2

Page 19: The Design of a Micro-turbogenerator

Figure 1.1: A micro-turbine engine operating a Brayton cycle, Spadaccini et al.(2003)

simple operation, high power density, and ability to use a wide range of fuels. And at

the micro-size level, a turbogenerator is an ideal choice also due to its theoretically

simple operation and low complexity; there are minimal moving parts, the compres-

sion, combustion, and expansion processes are steady state, and all the components

can be designed with planar geometry, allowing the utilization of micro-fabrication

techniques. In addition, when compared to its conventionally sized counterpart a

micro-heat engine has the advantage of benefiting from the cube-squared law, where

as the characteristic length of the engine decreases, its power density increases. This

can be shown as follows.

For a control volume, the output shaft power of a turbine is

9Wt ηt 9mcppTt4 Tt5q (1.1)

where ηt is the turbine efficiency, 9m is the mass flow, cp is the specific heat capacity at

constant pressure, Tt4 is the total temperature pre-turbine, and Tt5 is the total tem-

perature post-turbine. Factoring out Tt4 and using an isentropic expansion process,

this can be shown to be

9Wt 9mcpTt4p1 rPt5Pt4sγ1γ q (1.2)

where Pt represents the total pressure of the gas. As can be seen, for fixed conditions

such as efficiency, pressure, and temperature, the power output will scale linearly

3

Page 20: The Design of a Micro-turbogenerator

with the mass flow. And again for fixed conditions, the mass flow will scale linearly

with the area, or with the square of a characteristic length, L2. As a result, the

power output will scale with the square of the characteristic length.

9W 9 9m 9 A 9 L2 (1.3)

The volume of the device however will simply be proportional to characteristic length

of the system cubed, L3. Therefore for fixed conditions, the ratio of power to volume,

or the power density of the device will be

9W

V9 L2

L39 1

L(1.4)

Such a simple analysis shows the advantage of micro-sized gas turbines. If one were

to simply scale a gas turbine with an outer diameter of 1 meter to the size of a micro

gas turbine with an outer diameter of 1 cm and maintain its operating conditions

and efficiency, the power density of the device would increase 100 fold.

In order to extract electrical energy from a gas turbine however, an electrical

generator must be attached to its output shaft. It follows then that a similar analysis

must be presented to show the scaling laws that govern electrical generators. For

this work, a permanent magnet micro-generator was selected for reasons to be later

explained.

As shown by Faraday’s law of induction, the induced voltage in a loop of coil is

equal to the time rate of change of the magnetic flux in the area circumscribed by

the coil.

ε dΦ

dt(1.5)

where the flux is determined by

Φ » »

~B d ~A (1.6)

4

Page 21: The Design of a Micro-turbogenerator

As such, the induced voltage should be proportional to the area, or a characteristic

length squared, and the rotational speed of the magnetic field over the generator

coils.

ε 9 A Ω 9 L2 Ω (1.7)

As shown in the maximum power transfer theorem, for any given voltage, the maxi-

mum output power is related to the resistance of the generator coils by

9Wmax ε2

4 Rcoil

(1.8)

with the resistance of the coils being

R ρ LA

9 1

L(1.9)

Therefore, plugging in our voltage and resistance relationships, the power of the

generator system should scale as

9Wmax 9 L5 Ω2 (1.10)

However, with regard to turbomachinery, the tip speed of the turbine blades must

always be commensurate with the speed of the gases, which do not scale with the

size of the device. As such

Ω 9 1

L(1.11)

Inserting this conclusion into the power relationship shows that

9W 9 L3 (1.12)

As such, the power density of the magnetic generator should scale as

9W

V9 L3

L39 1 (1.13)

5

Page 22: The Design of a Micro-turbogenerator

and therefore no benefit would be seen with miniaturization. However, such a benefit

is observed in real world scenarios as a result of scaling laws that affect heat dissipa-

tion. As shown in figure 1.2, for a given current density, the heat creation will scale as

L3. However, conductive and convective heat dissipation will scale with the surface

area or L2, making the ratio of heat dissipation to heat creation scale as 1L

. As a

Figure 1.2: Scaling laws for heat dissipation in a conductor, Cugat et al. (2003)

result, larger current densities are acceptable in micro-sized generators than in their

conventionally sized counterparts. This allows generator power density to an extent

also scale with the inverse of the characteristic length of the system at the expense

of efficiency. This has been confirmed in permanent magnet micro-generator tests

by Arnold et al. (2006b), where a power density of 59MW m3 was reported, whereas

typical macro-sized generators typically achieve around 20MW m3. Alternatively,

with size reduction, volume that was previously dedicated to cooling elements can

instead be dedicated to power producing elements, allowing the power density to

increase while maintaining efficiency.

As demonstrated, given equal non-scaling operating conditions such as tip speed,

efficiency, pressure ratio, etc., the power density of both the gas turbine and perma-

nent magnet generator scale indirectly proportional to the characteristic length of

the turbogenerator system. However, as shown in this thesis, the efficiencies of these

components typically do not scale favorably with size, and thus the power density

does not scale quite as well as this simple analysis would suggest. Regardless, the

power density does scale well enough for micro-turbogenerators to provide sufficient

amounts of power for many applications. Likewise, given sufficient components effi-

6

Page 23: The Design of a Micro-turbogenerator

ciencies, the utilization of hydrocarbon fuels enables these devices to achieve energy

densities many times greater than conventional batteries. As shown in figure 1.3,

using ethanol as fuel, a thermal efficiency of roughly 13% and a generator efficiency

of 80% results in a device whose energy density can reach five times that of current

lithium-ion batteries. Thus there are many potential advantages associated with

Figure 1.3: Maximum energy density of a micro-turbogenerator relative to lithium-ion batteries as a function of thermal efficiency, generator efficiency, and fuel source

micro heat engines, particularly turbogenerators. They exhibit high energy density

relative to batteries due to their fuel source. They also possess sufficiently high

power densities due to scaling laws to power mobile devices and should theoretically

exhibit higher power densities than commercial fuel cells. For these reasons, micro-

turbogenerators are ideal power sources to run certain classes of devices, in particular

mobile devices that require endurant power sources and have relatively large power

requirements such as robotics, uav’s, and high power portable electronics. There

exists one other potential advantage associated with micro-turbogenerators. Due

to the power density scaling laws, if the devices are fabricated cheaply using large

volume manufacturing techniques as in the semiconductor industry, multiple micro-

7

Page 24: The Design of a Micro-turbogenerator

turbogenerators could be used in parallel to provide a large power output at a fraction

of the size of an equivalent macro sized machine.

1.1.2 Challenges

Despite all the advantages that could be attributed to micro-turbogenerators, a self

sustaining engine has never been designed and run due to the many difficulties and

challenges associated with their design, fabrication, and operation.

Gas turbines often utilize a Brayton cycle, where ideally gases are isentropi-

cally compressed, then heated reversibly, and then isentropically expanded through

a turbine. As can be seen in figure 1.4, due to heat addition, the change in total

temperature across a compressor and turbine with the same pressure ratio diverges.

This allows a gas turbine to not only create its own pressurized gas source, but also

deliver excess shaft power to a generator to create electric power. In real systems

however, all of these processes are done in an irreversible entropic manner. The

differences between the reversible and non-reversible cycle paths can also be seen in

figure 1.4. For both the compressor and turbine portions of the cycle, the amount of

Figure 1.4: Reversible and non-reversible TS diagrams for a Brayton cycle

8

Page 25: The Design of a Micro-turbogenerator

energy transfer will be

9W 9m cp ∆Tt (1.14)

As can be seen in figure 1.4 however, for any given pressure ratio, ∆Tt is larger for

the compressor and smaller for the turbine in the non-isentropic case relative to the

isentropic case. The result of this is that the compressor requires more input power

to achieve its pressure ratio, and the turbine delivers less power output for the same

pressure ratio. And because the compressor receives its shaft power from the turbine,

there then exists some efficiency below which the Brayton cycle will no longer close

as a result of the compressor requiring more power than the turbine can provide at

a given pressure ratio. This can be shown as follows. As stated before, the power

output for a turbine is

9Wt ηt 9mcpTt4p1 rPt5Pt4sγ1γ q (1.15)

Likewise, the power input for a compressor is

9Wc ηc 9mcpTt2prPt3Pt2sγ1γ 1q (1.16)

Equating these two until the power requirements are equal results in

ηt ηc ¥ Tt2prPt3Pt2sγ1γ 1q

Tt4p1 rPt5Pt4sγ1γ q

(1.17)

and therefore, for any given pressure and temperature ratios, there are minimum

component efficiencies that are required for the turbine to supply sufficient shaft

power to the compressor. If the efficiencies are below this threshold, the turbine

cannot power the compressor, and therefore no high pressure gas will be available to

power the turbine, and the cycle will not close.

For micro-turbomachinery unfortunately, the typical sources of entropic losses are

much greater than in their conventional counterparts due to relatively higher viscous

9

Page 26: The Design of a Micro-turbogenerator

effects and manufacturing constraints. Some of these turbomachinery loss sources

are tip clearance losses, trailing edge mixing losses, end wall losses, and viscous losses

in the boundary layer.

In addition, there are many forms of losses that are practically unique to micro-

turbine engines. For example, due to the planar geometry dictated by MEMS fabri-

cation constraints, there exist many right angle turns in the flow path of the working

fluid as shown in figure 1.5, which lower the total pressure of the fluid. Another issue

Figure 1.5: Gas flow path for a micro-turbine, Frechette et al. (2005)

that affects micro-turbomachinery especially is residual swirl leftover in the wake of

the turbine. As can be seen by analyzing the velocity triangles in figure 1.6, there ex-

ists some rotational speed (or tip speed) for the turbine blades where the tangential

velocity of the exit flow is zero, with the remaining flow having either only a radial

or axial component. This speed corresponds to the optimum speed of the turbine,

because all of the tangential kinetic energy of the gases has been transmitted to the

turbine shaft. However, as described in section 1.1.1, the turbine tip speed is inde-

pendent of size. As a result, extremely high values of RPM (often above 300,000)

are required for micro-turbomachinery, and this necessitates the use of high-speed

air bearings for the engine. Such bearing systems are extremely complex however

and have only recently shown sufficient progress to be reliably utilized as demon-

strated by Frechette et al. (2005). They also have yet to be correctly integrated

into a complete turbogenerator system. The highest observed rotational rate for a

fully integrated turbogenerator system utilizing gas bearings is currently only 40,000

10

Page 27: The Design of a Micro-turbogenerator

Figure 1.6: Velocity triangles of a turbine stage

RPM as demonstrated by Yen et al. (2008b).

Micro-turbine engines also typically suffer from issues associated with heat trans-

fer due to the small characteristic length of the devices. As an example, it is unfavor-

able for heat to transfer from the hot exhaust gases of the turbine to the gases within

or traveling to the compressor. This can be explained conceptually as follows. For a

turbomachine with an incompressible working fluid, the power transfer between the

fluid and the blades will be

9W 9m∆Ptρ

(1.18)

Conceptually then, one can think of a Brayton cycle as working across a fixed pres-

sure ratio, with the combustor existing solely to heat the temperature of the gas

and lower the density so that the turbine can extract more power during expansion

11

Page 28: The Design of a Micro-turbogenerator

than the compressor requires for compression. For this reason, if the gas within the

compressor or the gas in route to the compressor is heated, then for a given shaft

RPM determined by material constraints, a lower pressure ratio will be achieved and

result in lower system efficiency. In conventional sized machines, the heat conduc-

tion from the turbine section to the compressor is often negligible relative to the

mass flow. However, for micro-sized machines with small characteristic lengths, the

thermal resistance is sufficiently low that strong measures must be taken to limit

heat conduction. As can be seen in figure 1.7, structural changes such as adding

Figure 1.7: Comparison of a first generation design relative to a more recent designshowing the thermal resistance structures between the turbine and compressor, Lang(2009)

thin and long thermal barriers between the turbine and compressor sections must be

made. This figure contrasts an early generation design versus a more contemporary

design. However, this small scale heat transfer effect can be used advantageously

as well. As an example, a recuperator with a high effectiveness transferring heat

12

Page 29: The Design of a Micro-turbogenerator

from the exhaust gases to the pre-combustion gases can be made with a very small

form factor due to these scaling laws. For conventionally sized machines, the size

and weight penalties of these heat exchangers often outweigh their thermodynamic

benefits. Micro-sized heat engines however can use heat exchangers very effectively

to theoretically improve performance because heat exchangers are relatively much

smaller in micro-systems due to the scaling laws of heat transfer.

In addition to the efficiency problems associated with micro-turbomachinery,

there are many problems encountered by micro-generators due to their small size

as well. However, these problems are not as severe as those that affect micro-

turbomachinery components. The issues effecting permanent magnet micro-generators

will be discussed, because this type of generator is the one that is ultimately chosen.

One of the primary difficulties with micro-sized permanent magnet generators

is magnetizing the hard magnetic material. Ideally, there would be a square wave

magnetization pattern on the magnets alternating between north and south polarity

as shown in figure 1.10 (cm 0). However, due to leakage flux (figure 1.9) in the

magnetizing heads (figure 1.8) used to magnetize the material, trapezoidal magnetic

Figure 1.8: A magnetizing head used to permanently magnetize magnet mate-rial, Gilles et al. (2002)

patterns are typically encountered. This results in a decreased time rate of flux

13

Page 30: The Design of a Micro-turbogenerator

Figure 1.9: Leakage flux permeating incorrect sectors during magnetization, Gilleset al. (2002)

Figure 1.10: Trapezoidal magnetization profile of an annular permanent mag-net, Das et al. (2006)

transversing the generator coils and therefore smaller voltages. This in turn results

in a reduced efficiency for a fixed power output requirement and a fixed internal coil

resistance. A solution to this problem is to use discrete magnet pieces that can be

easily magnetized and to place them around the periphery of the generator as shown

in figure 1.11. This also helps solve the issue of balancing the generator rotor because

the magnets can be weighed and appropriately arranged around the periphery in

order to minimize imbalance. Unfortunately, this technique also necessitates the

14

Page 31: The Design of a Micro-turbogenerator

Figure 1.11: Discrete magnet pieces arranged peripherally around the generatorrotor, Herrault (2009)

use of a robust silicon frame to keep the magnet pieces situated. This is because

the magnetic material is no longer annular and therefore cannot resist radial forces

associated with centripetal motion. An example of such a silicon frame can be

seen in figure 1.12. Despite the frame however, using discrete magnet pieces will

Figure 1.12: A silicon frame for discrete peripherally arranged magnet pieces, Her-rault (2009)

always lower the maximum achievable RPM relative to an annular magnet because

15

Page 32: The Design of a Micro-turbogenerator

an annular magnet is able to support some ring strain by itself. The result of lowering

the maximum RPM is the lowering of the generator output voltage. This negatively

affects power output and efficiency.

There are also many fabrication constraints that negatively affect performance.

For example, due to the limited number of materials which can be electrically de-

posited using current MEMS techniques, materials that don’t possess optimal char-

acteristics are often chosen. A good demonstration can be seen with Yen et al.

(2008a), where a nickel-iron alloy with a low saturation density of only 0.8T (versus

the standard Hiperco with a saturation value of 2.4T) was chosen as the back iron

material for their generator rotor. In addition, due to a MEMS depositing thickness

limitation, not only was the choice of material compromised, but so was the geome-

try, with the actual back iron being only 20% of the ideal thickness. This decreased

the effective permeability of the back iron and therefore lowered system voltage and

efficiency.

Permanent magnet generator performance is also negatively affected by rotor

stability constraints imposed by the gas bearings. The journal bearings typically have

specific damping and stabilizing performance values, and higher rotor weights and

angular velocity values cause the rotors to cross this threshold. As a result, generator

rotors are often constructed with low aspect ratio (low thickness to diameter ratios)

magnets and magnetic back irons. As described in chapter 3, this decreases the flux

output of the generator and therefore lowers system voltage, power, and efficiency.

Fabrication constraints imposed by the 2D planar geometry also negatively af-

fect the internal resistance of the generator coils. As can be seen in figure 1.13,

complicated multi-plane coil patterns utilizing many electrical vias are required to

achieve multiple coil turns per pole. This has the effect of greatly increasing the in-

ternal resistance beyond what one would expect from standard area, resistivity, and

length calculations, and therefore lowers the efficiency of the system for any given

16

Page 33: The Design of a Micro-turbogenerator

Figure 1.13: Coil winding patterns for 2D micro-generators, Arnold et al. (2006b)

power output.

Micro-generators are also more negatively affected by eddy current and hysteresis

losses than conventionally sized machines. Hysteresis losses typically scale with the

frequency of the device, f , and eddy current losses scale with the frequency squared,

f 2. And as previously explained, the rotational speeds of micro-turbines must be

extremely high. This results in high electrical frequencies for the attached gener-

ator and therefore relatively high eddy current and hysteresis losses. In order to

reduce these losses, a few possible options can be taken. The first is to have a non-

conducting, non-magnetic stator iron so that eddy currents and hysteresis cannot

exist. This will typically have the effect however of increasing the reluctance of the

magnetic circuit and therefore lowering the generator output voltage and efficiency.

Another option is to laminate the stator back iron as shown in figure 1.14. This

will reduce the severity of the eddy current losses without greatly increasing the

reluctance of the magnetic circuit.

Another issue that has a great impact upon generator performance is the effect

that temperature has on the magnetic remanence of permanent magnets. As can be

17

Page 34: The Design of a Micro-turbogenerator

Figure 1.14: Non-conducting silicon laminations within the stator back iron toreduce eddy current losses, Herrault et al. (2010)

seen in figure 1.15, the remanence of magnetic materials typically degrades with tem-

Figure 1.15: Relationship between magnetization and temperature for rare earthcobalt materials, Campbell (1996)

perature. This is especially true in the case of neodymium boron material, which is

typically the highest performing material at room temperature. This thermal degra-

dation not only reduces the magnetic performance of the engine, but also completely

eliminates certain magnetic materials from consideration and forces designers to use

18

Page 35: The Design of a Micro-turbogenerator

alternative materials that are typically considered as inferior.

Lastly, of immense importance for micro-sized permanent magnet generators is

their power output capabilities. For any generator design operating at a fixed RPM,

there will be a maximum amount of mechanical shaft power that the generator

can convert to electrical power and dissipate into the surroundings as heat through

loss mechanisms. Should the generator be unable to utilize or dissipate its input

shaft power at its design RPM, it will accelerate beyond this speed and most likely

mechanically fail. For this reason, assuming a high value of efficiency is sought, it is

required that the maximum power transfer capabilities of the generator be high and

sufficient enough to convert the input shaft power into electric power. This presents

a challenge however because the power density of the turbomachinery scales much

more favorably with size than does the power density of micro-generators, making it

difficult for the generator to utilize the relatively large shaft power input at a high

efficiency.

Another key component of a heat-engine is the combustor as shown in figure 1.16.

This component also sees significant challenges due to the scaling laws that affect its

Figure 1.16: Schematic of a silicon wafer single-zone combustor, Spadaccini et al.(2003)

performance. As described by Spadaccini et al. (2003), there is a length of time that

is required for most of the injected fuel to combust, τreaction, which is related to the

19

Page 36: The Design of a Micro-turbogenerator

chemical reaction rate by an Arrhenius equation.

τreaction rfuelsA rfuelsa rO2sb eKaRT (1.19)

There is also a length of time known as the residence time, τresidence, which is the

length of time that the gas mixture requires to travel through the combustor volume.

τresidence volume

volumetric flow rate V ρ

9m V P

9m R T(1.20)

The ratio of these two values is known as the Damkohler number.

Dah τresidenceτreaction

(1.21)

In order to combust all of the injected fuel and achieve a high combustion efficiency,

the Damkohler number must be greater than one. This is difficult to achieve with

micro-combustors because, due to the size of device, τresidence is quite low. For

fixed operating conditions such as pressure, temperature, and combustor volume,

the residence time can be increased by reducing the mass flow rate through the

device. This however will reduce the power density of the device, making it a non-

viable option. The other option is to dramatically increase the reaction rate which

can be done by one of two methods: increasing the combustion temperature or using

catalytic surface material. Increasing the combustion temperature can be achieved

by burning closer to stoichmetric conditions (in typical gas turbines, combustion

takes place far from this point). The gases must still be cooled before entering the

turbine however, and therefore this setup requires the use of a dual-zone combustor,

as shown in figure 1.17. The first zone is utilized to burn the fuel quickly, at high

temperature, and closer to stoichiometric conditions. The second zone is used to mix

and cool the gases. The combustion reaction rate can also be increased by utilizing a

catalytic surface as shown in figure 1.18. This insert works by lowering the activation

20

Page 37: The Design of a Micro-turbogenerator

Figure 1.17: Schematic of a dual zone combustor, Spadaccini et al. (2003)

Figure 1.18: A silicon micro-combustor with a platinum plated catalytic in-sert, Spadaccini et al. (2002)

energy of the reaction and thereby increasing the reaction rate.

As shown, micro-turbogenerators are extremely complex and there are many chal-

lenges associated with their operation. Each of the components, the turbomachinery,

combustor, and electric generator all face severe hurdles due to manufacturing lim-

itations and scaling laws that govern their performance. However, significant work

has been done towards developing these components and integrating them into a

working engine.

21

Page 38: The Design of a Micro-turbogenerator

1.1.3 Review of Previous Work

A substantial amount of work has been done towards developing all the components

of a micro-turbogenerator as shown by Lang (2009). Most of this work was done

independently however, with the components being meant to function with external

support as opposed to working in an fully integrated engine.

Significant hurdles have already been overcome with gas bearings and micro-

turbomachinery. Peirs et al. (2003) demonstrated an axial flow impulse turbine

with an outer diameter of 10 mm generating 28 W of shaft power at 160,000 RPM

(figure 1.19). Savoulides et al. (2008) fabricated and tested a silicon MEMS tur-

Figure 1.19: The 10 mm axial impulse turbine tested by Peirs et al. (2003)

bocharger up to an RPM of 480,000 (figure 1.20). The turbocharger was powered by

an integrated micro-turbine and created a pressure ratio of 1.21. All of the turboma-

chinery components were supported by hydrostatic thrust and journal gas bearings.

Similarly, Frechette et al. (2005) fabricated and tested a single stage 4.2 mm diameter

radial inflow MEMS turbine (figures 1.21 and 1.22). The turbine was supported by

hydrostatic thrust and journal bearings and reached speeds greater than 1,000,000

RPM with a corresponding shaft power of 5 W. Lastly, Lee et al. (2008) demonstrated

22

Page 39: The Design of a Micro-turbogenerator

Figure 1.20: SEM photograph of the MEMS turbocharger, Savoulides et al. (2008)

Figure 1.21: Cross sectional view of the turbomachinery, Frechette et al. (2005)

Figure 1.22: A bird’s eye SEM photograph of the MEMS turbine, Frechette et al.(2005)

23

Page 40: The Design of a Micro-turbogenerator

a multistage radial inflow turbine operating at 330,000 RPM and with a mechanical

shaft power output of 0.38 W.

Figure 1.23: A split SEM view showing the concentric turbine stages, Lee et al.(2008)

A great deal of effort has also been successfully dedicated to micro-combustor

research. Mehra et al. (2002) designed and tested a hydrogen gas fueled micro-

combustor (figure 1.24). The device demonstrated stable operation in a silicon

structure with a flame temperature in excess of 1600 K. Subsequently, Spadaccini

et al. (2003) designed a dual zone combustor with the intent of increasing fuel burn

efficiency and allowing hydrocarbon combustion (figure 1.25) . Exit gas temper-

atures were as high as 1800 K and the power density of the device corresponded

to 1100MW m3. In order to increase combustion efficiency of hydrocarbon fuels,

platinum plated catalyst inserts were tested in micro-combustor experiments (fig-

ure 1.26). The insert allowed the combustion of hydrocarbon fuels in this device

which was not possible without the catalytic inserts. Combustion efficiencies in ex-

cess of 40% were demonstrated with the insert.

Electrical generator technology has also extensively progressed over this time pe-

24

Page 41: The Design of a Micro-turbogenerator

Figure 1.24: A fully packaged micro-combustor, Mehra et al. (2002)

riod. Holmes et al. (2005) showed a low pressure ratio integrated turbo-generator

(figure 1.27) achieving a power output of 1.1 mW at 30,000 RPM. The device oper-

ated with ball bearings and had discrete magnetic components embedded within the

turbine structure. Raisigel et al. (2006) created a generator that operated at 380,000

RPM with external support and created 5 W of power. The same device was able to

function without external support at 58,000 RPM and produce 14.6 mW of power.

Figure 1.25: Schematic of a dual zone combustor, Spadaccini et al. (2003)

25

Page 42: The Design of a Micro-turbogenerator

Figure 1.26: Schematic of a combustor with a catalyst insert, Spadaccini et al.(2002)

Figure 1.27: Schematic of a ball bearing supported turbo-generator, Holmes et al.(2005)

Das et al. (2006) and Arnold et al. (2006b) modeled, fabricated, and tested perma-

nent magnet generators up to 120,000 RPM with DC power outputs of 1.1 W. That

generator design was then further optimized by Arnold et al. (2006a) by improving

the winding pattern (figure 1.28) and operating at a higher RPM. The optimized

windings showed significantly reduced resistance values for the generator coils. This

resulted in a net electrical power output of 8 W at 305,000 RPM. Herrault et al.

(2008) then tested and characterized micro-generator performance at high tempera-

tures, confirming as expected that generator performance would decrease at elevated

temperatures. He demonstrated however that the generator performance can remain

adequate for micro-heat engine electrical power generation with careful selection of

materials (see figure 1.29). Herrault et al. (2010) then took the first steps to inte-

grate micro-generator technology into a silicon structure capable of supporting gas

26

Page 43: The Design of a Micro-turbogenerator

Figure 1.28: A comparison of the original and optimized winding patterns, Arnoldet al. (2006a)

Figure 1.29: A comparison between SmCo and NdFeB generators as a function oftemperature, Herrault et al. (2008)

27

Page 44: The Design of a Micro-turbogenerator

bearings and micro-turbomachinery, a key requirement for creating a self-sustaining

micro-turbogenerator. His device also contained a laminated stator-iron to reduce

eddy current losses and increase electrical efficiency.

At Duke, research has been conducted into ejectors and injectors. These are static

devices which utilize vaporized gases from boilers to pump secondary fluids. Initial

work was done to show that an ethanol vapor powered ejector could create a sufficient

suction draft to replace a compressor for a micro-turbine or to power hydrostatic gas

bearings during the startup transient periods of operation, Gardner et al. (2010b).

Subsequent work created detailed loss models for ejector performance, Gardner et al.

(2010a). In particular, the relationship between power density and size as a result

of scaling effects were shown (figure 1.30) as was the relationship between suction

draft and mass flow (figure 1.31). This work is significant, as will be explained

Figure 1.30: Power density and ejector efficiency as a function of the nozzle throatdiameter, Gardner et al. (2010b)

in detail later, because it allows a micro-turbine to operate without a compressor

and the cycle closing disadvantages of a Brayton cycle and without the condensation

28

Page 45: The Design of a Micro-turbogenerator

Figure 1.31: Suction draft performance versus entrainment ratio, Gardner et al.(2010a)

disadvantages associated with a Rankine cycle.

As shown, many of the components which make up micro-turbogenerator systems

have been heavily studied and researched. However, few studies have been conducted

towards integrating the various components into a working system. Frechette et al.

(2004) conducted high level design work for a micro-heat engine utilizing a Rankine

cycle (figure 1.32). His analysis showed an electrical power output of 1-12 W with

Figure 1.32: A schematic of a MEMS power plant utilizing a Rankine cy-cle, Frechette et al. (2004)

29

Page 46: The Design of a Micro-turbogenerator

an overall conversion efficiency of 1-10%. However, his device was never fabricated.

Yen et al. (2008a) designed an integrated turbogenerator system that utilized a

pressurized gas source. His system was fabricated on a seven wafer thick silicon

stack. The permanent magnets were discrete pieces and were imbedded within the

silicon structure of the turbomachinery, which was supported by hydrostatic journal

and thrust bearings. For this device, the generator windings and their connectors

were integrated into the silicon structure. This was the first integrated system to

have an electric generator powered by micro-turbomachinery and to have all the

components supported by gas bearings. A schematic of the device can be seen in

figure 1.33. It was designed for a rotational speed of 360,000 RPM and a power

Figure 1.33: Cross sectional schematic of the fully integrated permanent magnetturbogenerator, Yen et al. (2008b)

output of 1.5 W. However, due to a structural oversight, it was only able to accept

sufficient pressure to reach 40,000 RPM, resulting in a power output of 19 mW.

Regardless, this is arguably the most advanced silicon micro-turbogenerator to date.

As shown in more detail in chapter 4, Camacho et al. (2010) integrated two micro

turbines with a permanent magnet generator, power electronics, and an ejector. The

system was unique in that it lacked a compressor, but rather used an ejector to create

the pressure gradient to drive the turbine. The ejector was powered by combustion

30

Page 47: The Design of a Micro-turbogenerator

from an ethanol boiler, making this the first micro-engine to convert the heat from

combustion into electricity. The first of the two turbogenerators can be seen in

figure 1.34.

Figure 1.34: The original turbogenerator with the protruding generator leads con-nected to the power electronics board

As shown, a great deal of work has gone into developing the subsystems and

components for a micro-turbogenerator. The knowledge from this research is only

just now starting to be integrated together to form functioning engines. As such,

should there be a strong economical argument for micro-turbogenerators, a great deal

of useful information exists in the scientific literature from which one could begin to

design a working system.

1.2 Research Objectives, Goals, and Design Overview

The goal of this thesis is to achieve a few specific tasks that are required to design a

micro-turbogenerator and accurately model its performance. The first task is to show

the relation between efficiency and various turbine parameters such as reaction, stage

count, RPM, and size. This is of critical importance because it allows the designer

31

Page 48: The Design of a Micro-turbogenerator

to predict the actual mechanical shaft power output for given operating conditions

such as inlet pressure and temperature. It is also especially important in the case of

any Brayton cycle, because as described in 1.1.2, this efficiency value is necessary to

determine if the cycle can close for given operating conditions. The second task is to

accurately model micro-generator performance. This is important because it allows

the designer to predict the total system efficiency, from fuel burn to electric power

output, of a micro-turbogenerator. The total system efficiency is a crucial value in

determining whether a micro-engine would be superior to a fuel cell or battery for any

given application. It is also important to properly model the generator performance

so that turbomachinery and generator components can be correctly matched. If they

are designed independently and not correctly matched, then either the generator,

the turbomachinery or both will be working off design, with detrimental effects on

performance. Lastly, it will be demonstrated that, with no moving parts, an ejector

can be utilized to not only start-up a micro-turbogenerator through the transient

startup phase, but to also drive it in steady state and create electrical power, thus

presenting a relatively easy means to convert the heat from combustion into electric

power.

32

Page 49: The Design of a Micro-turbogenerator

2

Micro-turbine Design and Performance

2.1 Concepts of Turbomachinery

2.1.1 Operating Concepts

A turbomachine functions by absorbing or imparting energy, more specifically total

enthalpy, to the fluid traveling through it. A diagram of a general turbomachine can

be seen in figure 2.1. As it is a rotating machine, the power transfer is simply the

Figure 2.1: A control volume for a general turbo-machine, Mattingly et al. (2006)

33

Page 50: The Design of a Micro-turbogenerator

torque times the rotational velocity.

9Wt τ ω (2.1)

The torque on the blades of the turbomachine can be determined from the change

in angular velocity between the fluid leaving and exiting the device.

τ 9mpr1Vu1 r2 Vu2q (2.2)

The equation for power can then be re-written as

9Wt 9mωpr1Vu1 r2Vu2q (2.3)

This equation is known as the Euler turbine equation. It is useful however to expand

this equation in a way that sheds light on the various components of energy transfer.

Defining the product of ω and r as tip speed, U , we see that

9Wt 9mpU1Vu1 U2Vu2q (2.4)

Observing the velocity triangle in figure 2.2, we see that the flow leaving a turbine

Figure 2.2: Velocity triangles for the exit flow from a rotor blade

34

Page 51: The Design of a Micro-turbogenerator

blade will have some absolute exit velocity, V2, that is made up of its axial and

tangential components, Vm2 and Vu2. Note that the blade velocity U2, relates the

absolute velocity V2 to its relative velocity Vr2. Using geometry, we see that

V 2m2 V 2

2 V 2u2 (2.5)

V 2m2 V 2

r2 pU2 Vu2q2 (2.6)

Equating these two expressions results in

V 22 V 2

u2 V 2r2 pU2 Vu2q2 (2.7)

This can be further expanded and then simplified

V 22 V 2

u2 V 2r2 U2

2 2U2Vu2 V 2u2 (2.8)

U2Vu2 1

2pV 2

2 U22 V 2

r2q (2.9)

A similar process can be shown for the blade inlet and results in

U1Vu1 1

2pV 2

1 U21 V 2

r1q (2.10)

Inserting these values into the Euler turbine equation, eq. 2.3, results in

9Wt 9m1

2

pV 21 V 2

2 q pU21 U2

2 q pV 2r2 V 2

r1q

(2.11)

While the Euler equation is most often used in calculations, this form of the equation

is superior at showing the underlying mechanisms of energy transfer between the

rotor and the fluid. It is worthwhile to discuss which energy transfer mechanisms

the different terms correspond to. The first term, V 21 V 2

2 , is the easiest term to

understand, as it directly corresponds to the absorption of kinetic energy by the

rotor blades. The second term, U21 U2

2 requires more explanation. As described

by Shepherd (1956), if one were to observe a fluid element in a rotating environment

35

Page 52: The Design of a Micro-turbogenerator

Figure 2.3: A fluid element in a centrifugal field

as shown in figure 2.3, then clearly some type of force is being applied to the element

towards the center. In this case, the only force available is pressure. In order to

maintain a fluid element in rotation about the center of rotation, a centrifugal force

of dmω2r must be present, where dm drdAρ. This centrifugal force is a result of

a pressure gradient across the fluid element in the radial direction, dPdA. Equating

the two forces results in

dPdA drdAρω2r (2.12)

Rearrangement gives

dP

ρ ω2rdr (2.13)

Integrating both sides of the equation gives

»dP

ρ»vdP 1

2ω2pr2

2 r21q

1

2pU2

2 U21 q (2.14)

From the combination of the first and second laws of thermodynamics and for an

adiabatic process, the left hand side of the above equation is equal to the change

in enthalpy, and thus shaft work done to the fluid. As shown, this is equal to the

term 12pU2

2 U21 q, and therefore this term corresponds to the work done to the fluid

for traveling from one radius to another in a centrifugal pressure field. The last

36

Page 53: The Design of a Micro-turbogenerator

term, 12pV 2

r2V 2r1q, corresponds to changes in the relative fluid velocity, which simply

corresponds to changes in static pressure and therefore enthalpy.

As shown, a turbomachine works by changing the kinetic energy of the working

fluid, changing its static pressure through centrifugal pressure fields, and changing

the static pressure by accelerating the fluid flow in the relative frame. The sign of

these value changes determines whether or not enthalpy is being added to the flow

or being extracted, that is whether or not the machine is working as a turbine or

a compressor. Turbomachines can be further classified by the extent to which they

utilize these different energy transfer mechanisms. In particular, the ratio of energy

transfer via static pressure changes to the overall energy transfer is an important

feature called reaction, which as will be shown, is a key parameter when classifying

turbomachine operation.

2.1.2 Conservation Laws and Governing Equations

For turbomachine design, the basic equations which govern fluid flow must be used,

that is the conservation of mass, energy, and momentum, a state equation, the isen-

tropic relations, and the equation for the speed of sound in a gas. These will be

quickly reviewed.

The conservation of mass simply states that for any control volume, which typ-

ically coincides with a blade row, the mass into the control volume must equal the

mass exiting the control volume

9m ρV A const. (2.15)

The conservation of energy in a control volume simply states that for a control volume

in steady state, the rate of energy transfer out of the control volume must equal the

rate of energy transfer into the control volume plus the rate of heat addition and the

37

Page 54: The Design of a Micro-turbogenerator

rate of work done to the control volume.

»ρpe u2

2q~V ~dA dQ

dt dW

dt(2.16)

However, because work can be done at the system boundaries by the pressure forces

of the fluid itself,³p~V ~dA, it is often useful to consider this separately in the energy

equation so that shaft work and pressure work at the boundaries can be separated.

Doing this and combining terms results in

»ρpe p

ρ u2

2q~V ~dA dQ

dt dWshaft

dt(2.17)

The left hand term is seen to be the total enthalpy of the fluid. Using this variable

instead, the scalar form of this equation on a per unit mass basis becomes.

ht2 ht1 9q 9wshaft cppTt2 Tt1q (2.18)

The equation of state is the perfect gas law.

P ρRT (2.19)

The isentropic equations show the relation between changes in temperature and

pressure

Tt2Tt1

pt2

pt1

γ1γ

(2.20)

Using the speed of sound, these isentropic relations can show useful relationships

between total and static values and Mach number.

TtT 1 γ 1

2M2 pt

p

1 γ 1

2M2

γγ1

(2.21)

With the use of all these equations, a 1-D mean line analysis can be performed to

design isentropic turbomachinery.

38

Page 55: The Design of a Micro-turbogenerator

2.1.3 Turbine Reaction, Flow Type, and Stage Count

One of the primary ways to characterize a turbine is to define its reaction, which is

the ratio of energy transfer due to static pressure changes to overall energy transfer.

R rpU21 U2

2 q pV 2r2 V 2

r1qsrpV 2

1 V 22 q pU2

1 U22 q pV 2

r2 V 2r1qs

(2.22)

This ratio is very important because many turbomachinery properties such as design

RPM and efficiency are strongly correlated with reaction. The relationship between

efficiency and reaction will be shown in section 2.2.1, and the RPM relationship will

be shown here.

Imagine some jet stream flowing into a row of turbine blades at an angle of alpha

as shown in figure 2.4. Because the relative velocity Vr1 is equal to Vr2, the fluid was

Figure 2.4: Velocity triangles for an impulse turbine rotor blade row

neither accelerated nor decelerated in the blade row and therefore the static pressure

did not change overall. The absolute kinetic energy did change however, as noted by

the fact that magnitude of V2 is less than V1. Therefore, because none of the energy

transfer was associated with a static pressure change, the reaction of this turbine

row is zero. Now imagine that we wanted to drop the pressure within the blade row,

this would involve increasing the absolute value of Vr2 so that it was greater than

39

Page 56: The Design of a Micro-turbogenerator

Vr1. If we wanted to absorb the same exit kinetic energy however, such that V2 was

still the same as before, we see that the rotational speed U would have to increase

as shown in figure 2.5. This demonstrates then, that for absorbing the same amount

Figure 2.5: Velocity triangles for a reaction turbine rotor blade row

of kinetic energy from the incoming flow, a reaction turbine needs to operate at a

higher RPM than an impulse turbine. If it was not able to, this would result in

sufficient non-utilized kinetic energy remaining in the exit flow.

Often, the case is that material constraints or the bearing system limit the max-

imum allowable RPM of the turbine blades. As a result, significant kinetic energy

could remain in the flow in the form of tangential velocity or swirl. The solution to

this problem is to use multiple blade rows as shown in figure 2.6. The term for this

is multi-staging, and it allows you to absorb all tangential kinetic energy in the flow.

As will be discussed though, this is not without a cost, as it increases the production

of entropy within the turbine blades, with corresponding losses in efficiency.

Turbines also have different types of flow configurations and are named radial,

mixed, or axial. For a radial turbine as shown in figure 2.7, the fluid flows in the

direction perpendicular to the axis of rotation. Likewise, for an axial turbine, the fluid

flows in the direction parallel to the axis of rotation. In a mixed flow machine, the

flow enters axially and then leaves radially, or vice versa. Due to the manufacturing

process constraints for micro-turbines, their configuration is often limited to a purely

40

Page 57: The Design of a Micro-turbogenerator

Figure 2.6: Multistage turbine arrangement

Figure 2.7: A radial micro-turbine, Lang (2009)

41

Page 58: The Design of a Micro-turbogenerator

radial design when constructed out of silicon.

2.2 Micro-turbine Design

2.2.1 Turbine Efficiency and Loss Accounting

As described in section 1.1.2, an important issue with turbomachinery with regard

to closing thermodynamic cycles is efficiency. For a turbine dedicated to converting

enthalpy into shaft work, the efficiency can be define as

η 9Wt,realized

9Wt,potential

(2.23)

where the potential work extraction rate has already been shown to be

9Wt,potential 9m1

2

pV 21 V 2

2 q pU21 U2

2 q pV 2r2 V 2

r1q

(2.24)

As mentioned in previous sections, a simple way to lower the net work from a turbine

is to leave swirl velocity in the exit flow due to insufficient angular velocity. This

swirl velocity represents wasted kinetic energy that could have been extracted by the

turbine had it been operating at its optimal speed. Assuming completely isentropic

flow, the term for the ratio of actual energy absorption to ideal absorption is known

as effectiveness.

ε rpV 21 V 2

2 q pU21 U2

2 q pV 2r2 V 2

r1qsrV 2

1 pU21 U2

2 q pV 2r2 V 2

r1qs(2.25)

The effectiveness can be derived analytically for reaction values of R=0 and R=1/2

as follows. For an impulse turbine, because it derives all of its energy from kinetic

energy, its effectiveness ratio is simply

ε U∆Vu12V 2

1

(2.26)

42

Page 59: The Design of a Micro-turbogenerator

Referring again to the velocity triangles for the impulse case in figure 2.4, we see

that

Vu1 V1cospα1q and Vu2 V2cospα2q

So

∆Vu V1cospα1q pV2cospα2qq

We know from relative velocities however that

V1cospα1q Vr1cospβ2q U and V2cospα2q V2rcospβ2q U

Plugging these values into the above equation results in

∆Vu Vr1cospβ1q U pV2rcospβ2 Uq Vr1cospβ1q pV2rcospβ2q

Because this is an impulse turbine by definition, the relative inlet and exit velocities

must be equal and therefore

Vr1 Vr2 and β1 β2

Plugging these values into the above equation results in

∆Vu 2Vr1cospβ1q 2 rV1cospα1q U s (2.27)

Plugging this into the impulse effectiveness equation, eq. 2.26, results in

ε 2U rV1cospα1q U s12V 2

1

4U

V1

cospα1q 4

U

V1

2

(2.28)

As can be seen by this equation, a key parameter in determining the effectiveness of

a turbine is the ratio of UV1 which is known as the velocity ratio.

The same procedure can be done for the case of a reaction turbine assuming that

V 20 V 2

1 and using the defining relations of a 50% reaction turbine.

V1 Vr2 V2 Vr1 α1 β2 α2 β1

43

Page 60: The Design of a Micro-turbogenerator

The end result is

ε 2pUV1

qcospα1q U

V1

2

(2.29)

These plots, as well as those for a two stage and three stage impulse turbine, can be

seen in figure 2.8, and demonstrate two key features. The first is, as was said before,

Figure 2.8: Effectiveness of impulse turbines and a reaction turbine as a functionof the velocity ratio for α1 20

that in order to achieve the same effectiveness for a fixed value of V1, a reaction

turbine must operate at a higher RPM, and single stage turbines must operate at a

higher RPM that multi-stage turbines. The second conclusion is that there is a value

of UV1

which maximizes the effectiveness, and therefore for fixed operating conditions,

there is a specific value of RPM which maximizes effectiveness. It should be noted

that while reaction turbines require higher values of RPM to operate effectively, their

advantage lies in increased efficiencies. This is a result of two principle features. The

first is that, in contrast to an impulse turbine, there is a favorable pressure gradient

throughout the entire device (NGV’s, rotors, and stators). As will be shown, this

44

Page 61: The Design of a Micro-turbogenerator

reduces viscous losses in the boundary layers. The second is that for the same

pressure ratio, the average cubed velocity value of the gas stream (proportional to

the rate of energy loss) in a reaction turbine is less than in a impulse turbine. This

is a result of the manner in which reaction turbines and impulse turbines expand

their gases. For an impulse turbine, all of the pressure is converted to dynamic head

within the NGV’s. This results in very high gas speeds at the NGV exit and within

the first few rotors and stators before this kinetic energy is absorbed. For a reaction

turbine, the pressure is converted into dynamic head at a gradual rate over the length

of the device, and this kinetic energy is constantly being absorbed by the rotors. The

result is that, for the same overall pressure ratio, extremely high gas speeds are not

observed in reaction turbines as they are in impulse turbines. Therefore, the average

cubed value of velocity is less in reaction turbines, resulting in less energy dissipation.

As shown, the losses associated with effectiveness are unrelated to irreversibility and

entropy creation, they are simply a result of leaving kinetic energy in the flow in the

form of swirl. Other loss mechanisms in the flow that are irreversible and lower the

power output of the turbine will now be discussed.

For internal flow through turbomachines, it is not common to use drag as a metric

of loss as it is in aircraft. This is because it is difficult to define the direction in which

the drag vector acts in a turbomachine, Denton (1993). A more a common metric

of loss in turbomachines is entropy creation, which in a stationary adiabatic blade

row is a measure of the total pressure loss.

∆s R lnpPo2Po1q (2.30)

The advantage of using entropy generation as the metric for loss is that entropy

does not depend on the reference frame from which it is measured. The same values

of entropy will be measured whether the flow is viewed from stationary or rotating

blade rows.

45

Page 62: The Design of a Micro-turbogenerator

As stated by Greitzer et al. (2004), in a non-ideal flow with no heat exchange or

shaft work, the amount of work required to bring an irreversible flow process back

to its initial state is related to entropy creation by

wrev Tt∆s (2.31)

This equation can be reasoned out quite simply. Imagine an irreversible and adiabatic

flow process through a throttle device as shown in figure 2.9. Because the throttle is

Figure 2.9: An irreversible flow process through a throttle

adiabatic, there is no heat addition to the flow. Likewise, because there are no moving

parts within the throttle, no shaft work is being performed on the fluid. Referring

back to the energy equation, eq. 2.18, this means that the stagnation enthalpy must

be constant as well as the internal energy at stagnation for the inlet and outlet flows

(which is only a function of temperature for a perfect gas). From the first law of

thermodynamics however, if the stagnation internal energy has remained constant,

then

q w (2.32)

It is also known that for reversible heat addition

dqrev Tds

In the abstract then, one can think of an irreversible entropy creating process as one

that is adding heat to itself. From equation 2.32 however, this can only be done

if work was performed by the fluid. Combining these two thought processes, an

46

Page 63: The Design of a Micro-turbogenerator

irreversible process can be thought of as a fluid taking its ability to perform work

and converting it into heat in a manner which cannot be undone, reducing the fluids

ability to perform useful work on its surroundings. And just as heat transfer can be

determined by measuring the entropy change, so too can this internal conversion of

energy.

9wlost T 9ds (2.33)

In order to determine the efficiency of a fluid system then, one must keep track of

all the sources of entropy creation in a flow process, relate this entropy creation to

lost work, and compare this value to the fluids initial ability to perform work. This

requires the ability to quantify entropy creation for all the irreversible processes in

turbomachine flow such as boundary layers profile losses, fluid mixing losses, and tip

clearance losses. Before these losses can be calculated though, many parameters of

the flow must be characterized.

2.2.2 Characterization of the Flow

For all three loss mechanisms modeled in this work, various parameters of the flow

over the blades must be known. For example, the velocity distribution must be

estimated over each blade in order to estimate both the surface pressures and the

boundary layer parameters such as displacement thickness and momentum thickness.

Following Denton (1993), the difference in velocities on the blade surfaces can be

roughly estimated by assuming that the blade loading is constant. By then using

the definition of blade circulation and setting it equal to zero for the non-lifting flow

channel, it can be seen that

Vs Vp p

CVxptanα2 tanα1q (2.34)

Using Vr as a reference value, idealized velocity distributions over the blades can then

be assumed as shown in figure 2.10 for the different turbine types. With a knowledge

47

Page 64: The Design of a Micro-turbogenerator

Figure 2.10: Simplified velocity triangles for idealized impulse and reaction blades

of the velocities over the blades, the boundary layers can be analyzed to determine

useful parameters such as the momentum and displacement thicknesses.

In the case of laminar flow, the Falkner-Skan velocity profiles were solved for

each pressure gradient encountered as shown by Cebeci and Bradshaw (1977). For

turbulent flow, the following equations from Greitzer et al. (2004) were used for

displacement thickness and momentum thickness.

δpxq 0.37x

Re15x

(2.35)

θpxq 0.036x

Re15x

(2.36)

The same equations were used for both favorable and zero-pressure gradient flows,

as the state of the boundary layer was not determined to differ much between the

two, see White (1991)

2.2.3 Boundary Layer Losses

As described by Denton (1993), the entropic change due to profile loss in the bound-

ary layer is conveniently expressed non-dimensionally as the viscous dissipation co-

48

Page 65: The Design of a Micro-turbogenerator

efficient.

Cd T 9Saρ V 3 (2.37)

The value Cd is a function of the Reynolds number based off of momentum thickness.

This function changes for turbulent, laminar, favorable, and non-favorable boundary

layers. From Denton (1993), the dissipation coefficient for zero pressure gradient

turbulent flow was modeled as

Cd 0.0056 Re16θ (2.38)

However, simulations by Cebeci and Carr (1978) show that the dissipation coefficient

for accelerating turbulent boundary layers can be significantly less. This value was

modeled as shown below.

Cd 0.0205 Re0.417θ (2.39)

The Pohlhausen family of velocity profiles may be integrated to attain dissipation

coefficients for laminar boundary layers in accelerating, decelerating, or separated

flows in terms of the Pohlhausen parameter, λ, as shown by Schlichting (1968).

These results can be simplified into a laminar dissipation coefficient.

Cd β Re1θ (2.40)

The value of β can range from 0.220 for highly favorable pressure gradients to 0.173

for boundary layers with no pressure gradient. The graphs depicting the various

values of Cd for turbulent, laminar, favorable, and zero pressure gradients are plot-

ted in figure 2.11. As can be seen, the entropy generation in a laminar boundary

layer is significantly less than its turbulent value near the transition point. For this

reason, it is extremely important to accurately determine the state of the boundary

layer for real systems and to attempt to retain laminar flow for as long as possible.

For this analysis, a value of Reθ 200 was chosen as the transition point(Greitzer

49

Page 66: The Design of a Micro-turbogenerator

Figure 2.11: Dissipation coefficients for favorable,constant pressure, laminar, andturbulent boundary layers

et al. (2004)). Also, it can be seen that for turbulent flows, the dissipation coefficient

in an accelerating boundary layer is significantly less at higher Reynold’s numbers,

highlighting the advantage of reaction turbines. Lastly, it can be observed that for

very low Reynolds, such as those that characterize micro-turbine flows, the dissipa-

tion coefficient is very high, and thus one should expect large losses as a result of

this mechanism. After determining the value of Cd across a blade, the total entropy

generation per unit height within the boundary layer up to the trailing edge can

then be computed for both the suction and pressure surfaces as described by Denton

(1993).

9S ¸

Cs

» 1

0

Cd ρ V3

TdpxCsq (2.41)

This provides a mechanism to quantify the creation of entropy within the boundary

layers for the entire turbine and relate this to lost work potential. A similar analysis

can be done for lower and upper walls of the rotor and stator rows. And as can be

50

Page 67: The Design of a Micro-turbogenerator

seen in figure 2.7, micro-turbines typically have very large chord to height ratios as

a result of high centripetal stresses. This decreases the ratio of mass flow to wetted

area and therefore decreases efficiency.

2.2.4 Trailing Edge Losses

Entropy is also created in the wake of the blade trailing edges as shown in figure 2.12.

Three sources contribute to this entropy creation: the mixing out of the boundary

Figure 2.12: Wake mixing behind the trailing edge of two turbine blades

layers at the trailing edge, a lower base pressure on the blade tip than would exist in

an inviscid case( Roberts and Denton (1996)), and the combined blockage from the

boundary layers and the trailing edge itself. The last mechanism stated is indicative

of the other two and is very analogous to the losses in a dump diffuser. For this

reason, a quick analysis of the losses in a dump diffuser is warranted because its

sheds light on the actual loss mechanisms experienced at the trailing edge of the

turbine blades.

A dump diffuser, as shown in figure 2.13, is a device that expands the flow

abruptly as opposed to using slowly varying areas ratios as dictated by the isentropic

flow equations. This sudden expansion is an entropy creating process and therefore

lowers the total pressure of the fluid.

Creating a control volume within the dump diffuser as shown in the figure, a mass

and momentum balance can be done. From the conservation of mass and assuming

51

Page 68: The Design of a Micro-turbogenerator

Figure 2.13: A dump diffuser

an incompressible fluid we have

u1A1 u2A2 (2.42)

For the conservation of momentum, we need to know the pressure of the diffuser

back wall. It can be reasoned out, but also shown experimentally, that this pressure

is simply equal to the jet pressure. Noting that A2 A1 Awall, this results in

ρu1A1u1 p1A2 p2A2 ρu2A2u2 (2.43)

This can be rearranged to give

p1 p2 ρu22

A1

A2

ρu21

A1

A2

We can add dynamic head terms to both sides of the equation to determine the

total pressure loss and utilize mass flow to substitute out u2. Doing so and dividing

through by the inlet dynamic head results in

pt1 pt212ρu2

1

A21

A22

2A1

A2

1

52

Page 69: The Design of a Micro-turbogenerator

Defining A2

A1as the aspect ratio AR, this can be represented as

pt1 pt212ρu2

1

1 1

AR

2

The graph for this normalized total pressure drop can be seen in figure 2.14. Recall

Figure 2.14: Normalized total pressure drop in a dump diffuser as a function ofaspect ratio

from equation 2.30, that a drop in total pressure corresponds with an increase in

entropy, confirming that dump diffusion is an entropy creating process. As a side

note, notice that for infinitesimally small aspect ratios, the total pressure drop is

zero. Consequently, one can think of an isentropic diffuser as an infinite amount of

dump diffusers placed in series feeding into each other.

Looking back at the trailing edge of a turbine blade then, one can see how the

sudden expansion of the flow behind the trailing edge is identical to dump diffusion.

Therefore, the same exact method can be used to determine the losses as a result

of the trailing edge thickness and as a result of the displacement thickness due to

the boundary layer. A similar method can be performed to demonstrate why mixing

53

Page 70: The Design of a Micro-turbogenerator

out of the momentum boundary layer results in total pressure losses. This analysis

also clearly explains why a lower pressure on the trailing edge would result in total

pressure losses, because it would directly lower the static pressure at the mixed out

location.

Denton (1993) combines all three of these sources into a loss coefficient ,

ζ Cpb tw

wδ t

w

2

(2.44)

and from this we have an easy method by which to quantify the creation of entropy

from trailing edge mixing.

For micro-turbomachinery, these losses can be significant on account of the large

trailing edge thickness relative to the chord as shown in figure 2.15 due to feature res-

olution and manufacturing constraints (Lee et al. (2008)), high material stress levels

(Osipov (2008)), and the rate at which the displacement and momentum thickness

of the boundary layers grow over such small chord lengths.

2.2.5 Tip Clearance Losses

Entropy is also generated as the lower velocity, higher pressure gas from the pressure

side of the turbine blade spills over and mixes with the higher velocity gas stream of

the suction surface as shown in figure 2.16. As shown, this mechanism is analogous

to a small high velocity jet stream being injected into a larger and slower body of

fluid. As described in detail by Shapiro (1953), the entropy creation from such a

process is

ds

cp 2

VpVs

pk 1qM2

2

dm

m pk 1qM2dm

m(2.45)

By substituting these thermodynamic relationships

k cpcv

M2 V 2

kRTcp cv R

54

Page 71: The Design of a Micro-turbogenerator

Figure 2.15: Thick trailing edges of a micro-turbine as a result of feature resolutionand stress levels, Lee et al. (2008)

the equation can be rearranged to give

Tds 1

mV 2s p1

VpVsqdm (2.46)

We only need a few more pieces of information to utilize this equation. The first is

the velocity profiles that were shown previously. The profiles can be determined by

assuming constant blade loading and using the continuity equation. This results in

the following equations

Vs Vp p

CVxptanα2 tanα1q (2.47)

Vs Vp 2Vxcosα

(2.48)

The second piece of information required is the mass flow of the jet from the higher

pressure fluid traveling over the blade tip and into the suction flow. Using incom-

55

Page 72: The Design of a Micro-turbogenerator

Figure 2.16: A schematic of tip leakage over a turbine blade

pressible flow assumptions, this can be approximated as

dm Cdisg

d2ρ

1

2pV 2

s V 2p q

(2.49)

Cdis is the discharge coefficient that often must be determined experimentally. How-

ever, it was chosen as 0.6 for this analysis as that was seen as a typical observed

value in the literature. Combining all of these equations into a loss equation for tip

56

Page 73: The Design of a Micro-turbogenerator

clearance results in

T∆s CdisgC

V2hp cosα2

» 1

0

VsV2

21 Vp

Vs

gffe1

Vp

Vs

2dz

C(2.50)

Due to manufacturing tolerances, the ratio of tip gap to blade heights is often larger

for micro-turbines than it is for conventionally sized machines. For large scale gas

turbines, the values of tip clearance can be less than 1% of the blade height while

for micro-turbomachinery this value is often much greater (Jovanovic (2008)). As

a result, the losses associated with this mechanism in this analysis are also more

detrimental to micro-turbomachinery than they are to conventially sized machines.

2.2.6 Scaling Effects on the Efficiency of Gas Turbines

Using these equations, the entropy generation for each mechanism was calculated

and the associated power loss was numerically computed for one stage, two stage,

and three stage impulse turbines as well a single stage reaction turbine over a large

range of outer diameter and RPM. The turbines were modeled as axial turbines that

are scaled geometrically based on the outer diameter while retaining the same mass

flux, a total-to-static pressure ratio of Pt1Pe 1.85, and turbine inlet total temper-

ature of Tt1 1400 K. All impulse turbines were designed as velocity-compounded

Curtis turbines where all pressure drop occurs strictly in the first nozzle guide vane,

and subsequent stator rows only redirect the flow. The pitch to chord ratio was

numerically altered in order to minimize the combined losses for the rotors, stators,

and NGV’s. The results for this analysis are shown below.

Figures 2.17 - 2.20 show the effectiveness of the different turbines as a function of

their velocity ratio and diameter. Recall that effectiveness simply measures the ratio

of actual energy absorbed to the potential energy that could have been absorbed at

the optimum RPM. The flow is assumed to be isentropic and therefore lossless. For

57

Page 74: The Design of a Micro-turbogenerator

reference, lines of constant RPM have been drawn into the figures. A few expected

Figure 2.17: Turbine effectiveness as a function of diameter and the velocity ratiofor a single stage reaction turbine

features can be noted from these graphs. The first is that the optimum velocity

ratio required for the single stage reaction turbine is greater than for the single stage

impulse as expected. Likewise, the velocity ratios for the multistage turbines are

lower than their single stage brethren. The other observed feature is that for small

turbine diameters, extremely high values of RPM are required to reach the design

speed of the turbine. Note that even with isentropic flow, the effectiveness does not

reach 100% even at the optimal RPM. This is because there remains unutilized kinetic

energy in the purely axial exit flow. In order to achieve close to 100% effectiveness,

a diffuser would be required downstream of the turbine, allowing it to expand the

gases to sub-atmospheric conditions.

These graphs do not show the effect of the loss mechanisms however, which can

have large effects on overall efficiency. When loss mechanisms are included, the plots

change as shown in figures 2.21 - 2.24. Note that the effect of viscous drag in the

58

Page 75: The Design of a Micro-turbogenerator

Figure 2.18: Turbine effectiveness as a function of diameter and the velocity ratiofor a single stage impulse turbine

Figure 2.19: Turbine effectiveness as a function of diameter and the velocity ratiofor a two stage impulse turbine

59

Page 76: The Design of a Micro-turbogenerator

Figure 2.20: Turbine effectiveness as a function of diameter and the velocity ratiofor a three stage impulse turbine

Figure 2.21: Turbine efficiency as a function of diameter and the velocity ratio fora single stage reaction turbine

60

Page 77: The Design of a Micro-turbogenerator

Figure 2.22: Turbine efficiency as a function of diameter and the velocity ratio fora single stage impulse turbine

Figure 2.23: Turbine efficiency as a function of diameter and the velocity ratio fora two stage impulse turbine

61

Page 78: The Design of a Micro-turbogenerator

Figure 2.24: Turbine efficiency as a function of diameter and the velocity ratio fora three stage impulse turbine

direction of rotation is not accounted for. Therefore, at extremely small diameters,

the performance will be higher then depicted here, as the device begins acting more

as a Tesla turbine than a conventional bladed turbine.

From these figures, multiple observations can be made. The first is that, at

design speeds, reaction turbines show much greater efficiencies as small scales than

their impulse counterparts. This is for a few key reasons. The first is that for reaction

turbines, in the nozzle guide vanes, rotors, and stators, there is a continual favorable

pressure gradient. This favorable pressure gradient lowers viscous dissipation losses

in the boundary layer and reduces the growth rate of the boundary layer. Impulse

turbines have no pressure gradient after the NGV, and thus see large boundary

layer related losses. Another reason is that, assuming the same power output, since

a reaction turbine is rotating at a higher rate, it must be delivering less torque.

The corresponds to the blades being less loaded which implies that the pressure

62

Page 79: The Design of a Micro-turbogenerator

difference between the suction side and pressure side is lower in the reaction case

than in the impulse case. This will reduce losses due to tip clearance. The last

main reason that reaction turbines display higher efficiency is that the average value

of V 3 over the blade span, which is indicative of flow losses, is much lower. For

reaction turbines, static pressure is continually being converted into kinetic energy

as the blade passages converge. Thus the fluid is only traveling at a high relative

velocity towards the end of the blade passages. For an impulse turbine, all of the

available pressure is converted into dynamic head in the NGV’s, and therefore the

maximum velocity is seen throughout the rotor blade rows and stators. Because of

this, the average value of V 3 over the blade spans is higher in impulse turbines than

in reaction turbines, and therefore reaction turbines operate with lower losses.

Another interesting observation that can be made is that multi-staging is much

more effective for large diameters than it is for smaller diameters. This can be

reasoned out quite simply. For any turbine, it is extremely important to operate

at the design RPM in order to maximize effectiveness. If this cannot be done with

a single stage, more stages must be added. Adding stages however increases the

wetted area of the turbine blades and therefore increases viscous boundary layer

losses. Additional stages also increase the instances of trailing edge mixing losses and

tip clearance losses. For large turbomachines, these losses are not very significant

to begin with, and thus the benefit of adding multiple stages which allow on design

operation far outweigh the entropy creation penalties associated with the additional

stages. For smaller turbines however, these losses become much for significant and

a result, the benefits of multi-staging are attenuated by the increased viscous losses.

For this reason, it is of critical importance for micro-turbomachinery that the bearing

system allow the turbomachinery to have as few stages as possible and operate at

the design RPM.

Lastly, graphs can be shown depicting the power density of the various devices,

63

Page 80: The Design of a Micro-turbogenerator

operating at their design RPM, as a function of diameter. The results can be seen

in figures 2.25-2.28. These graphs display two key pieces of information. The first

Figure 2.25: Power density as a function of diameter for a single stage reactionturbine

Figure 2.26: Power density as a function of diameter for a single stage impulseturbine

is that as expected, the power density shows a linear relationship with characteristic

length over a large range of diameters. However, at very small length scales, the

64

Page 81: The Design of a Micro-turbogenerator

Figure 2.27: Power density as a function of diameter for a two stage impulseturbine

Figure 2.28: Power density as a function of diameter for a three stage impulseturbine

65

Page 82: The Design of a Micro-turbogenerator

increased entropic losses negate this effect and begin to level out the power density

and then even reduce it. The other conclusion, not surprisingly, is that the reac-

tion turbine maintains its linear power density relationship over a larger range of

turbine diameters. In addition, due to its overall higher efficiency and single stage

compactness, it possesses the highest power density of all the turbines analyzed.

A final observation can be made by looking at the alternate form of the Euler

turbine equation, eq. 2.11. As can be seen, there exists a mechanism to transfer

energy from the fluid to the rotor via the centrifugal pressure field. However, none of

the loss mechanisms analyzed here, which are the main forms of loss in a turboma-

chine, are related to this centrifugal pressure field. Therefore, energy transfer that

utilizes this mechanism is lossless. More on this is described by Cumpsty (1989). As

a result, for micro-turbomachinery, which is typically associated with high losses, as

large a fraction as possible of the total energy transfer should come from centrifugal

effects in order to increase efficiency. And because centrifugal effects are related to

static pressure changes and require high RPM, this means that the reaction should

be further increased, further highlighting the critical importance of the gas bearing

system.

2.3 Conclusion

In this chapter, loss models were presented for insufficient RPM exit losses, boundary

layer profile losses, tip clearance losses, and trailing edge mixing losses. In particular,

an analysis was presented for how these losses scaled with turbine size, type, and

configuration. The result was a series of graphs that showed efficiency values over

a large range of velocity ratios and diameters. A few key observations were made.

The first was that reaction turbines should be pursued as a result of their increased

efficiency and power density, which is due to their favorable pressure gradients. It

was also shown that the practice of multi-staging, while highly useful and effective

66

Page 83: The Design of a Micro-turbogenerator

at large values of diameter, is not as beneficial for smaller sized machines. It was

also confirmed that over a large range of diameter, the power density followed a

linear relationship as predicted by the scaling laws. At very small sizes, increased

viscous losses negated this effect however, resulting in a peak power density for each

configuration. Lastly, it was mentioned that radial turbomachines that use centrifu-

gal pressure fields for a significant portion of energy transfer should be utilized at

very small scales because an entire component of energy transfer is lossless. These

large centrifugal pressure fields require extremely high values of RPM however and

increase the reaction of the design. This further shows the importance of design-

ing effective gas bearings that can maintain the high RPM values required by high

reaction turbomachinery.

67

Page 84: The Design of a Micro-turbogenerator

3

Permanent Magnet Micro-generator Design andPerformance

3.1 Concepts of Permanent Magnet Generators

3.1.1 Operating Concepts, Scaling Laws, and Generator Selection

As described in Chapter 1.1.1, when a magnetic field varies within a contour in space,

by Faraday’s law of induction, a voltage will be induced around the path encircling

that contour proportional to time rate of change of the magnetic flux within the

contour.

ε dΦ

dt(3.1)

where the flux is determined by

Φ » »

~B d ~A (3.2)

The source of flux can come from different sources however. One solution is to use

electric currents in a wire such that a directed magnetic field is created as shown

in figure 3.1. The other solution is to use previously magnetized material, or a

permanent magnet, to create the magnetic field as shown in figure 3.2. In both

68

Page 85: The Design of a Micro-turbogenerator

Figure 3.1: A current conducting wire used to create a magnetic field, Nave (2011)

cases, the flux source is typically placed on a rotating structure such as a rotor so

that the magnetic field within an area is constantly changing, thereby creating an

alternating voltage source. If an electrical conducting material is then placed around

that contour path and then connected to a load, a current will flow and produce

electric power as shown in figure 3.3. This is the mechanism by which an electrical

generator works.

Typically however, a generator will have three sets of conducting coils arranged in

such a way that their voltages are out of phase by 120 as shown in figure 3.4. This is

termed three phase power. Three phase power is essential for an electrical generator

that receives its shaft power from a turbine because three phase power results in a

constant, non-sinusoidal, power draw from the generator, and therefore a constant

69

Page 86: The Design of a Micro-turbogenerator

Figure 3.2: A permanent magnet used to create a magnetic field, Nave (2011)

Figure 3.3: The induced voltage being used to create a current and produce work

power draw from the turbomachinery. This can be proven as follows. Imagine that

each phase of the generator created some sinusoidal voltage.

v1ptq Vpcospωtq v2ptq Vpcospωt 120q v3ptq Vpcospωt 240q

The resulting system power would then be

P ptq v21ptqR

v22ptqR

v23ptqR

Using the trigonometric identity cos2pαq r1 cosp2αqs2, this can be arranged to

70

Page 87: The Design of a Micro-turbogenerator

Figure 3.4: A three phase power system

give

P ptq 3V 2P

2R V 2

P

2Rrcosp2ωtq cosp2ωt 240q cosp2ωt 480qs

The term on the right always adds to zero for any value of t, and therefore the power

draw is constant at

Pt 3V 2P

2R

This is an essential property for a a device that is connected to turbomachinery, be-

cause turbomachines are designed to operate in steady state conditions. If a turbine

was connected to a single phase generator, then the turbine would see a sinusoidal

load and would therefore constantly be accelerating and decelerating and working

off-design, lowering its performance.

The other advantage with three phase systems is that it lets you arrange the

voltage coils in series as shown in figure 3.4. Notice that across leads A-C, the voltage

would be Vcc1 Va1a. The advantage of this is can most easily be seen in a phasor

diagram as shown in figure 3.5. As can be seen by trigonometry, the magnitude

of the combined voltage from both voltage sources is?

3 greater than the original.

This not only provides an easy manner through which the output voltage can be

increased, but it also increases the maximum power capabilities of the generator

and the efficiency relative to three independent single phase outputs. In many cases

however, an 3 phase alternating power source is not directly useful because many

71

Page 88: The Design of a Micro-turbogenerator

Figure 3.5: A phasor diagram for a wye connected generator

devices utilize DC power. In order to convert from AC to DC, a diode bridge rectifier

as shown in figure 3.6 must be used. A diode is a device that only allows current

Figure 3.6: A diode bridge rectifier

to flow in one direction. Taking this into account and observing the AC voltages in

figure 3.7, one can trace out the emitting and returning pathways for the current as

shown in figure 3.8. By noting that analogous pathways will be taken when each

one of the waveforms is near its AC peak, the observed rectified waveform across the

72

Page 89: The Design of a Micro-turbogenerator

Figure 3.7: The original AC voltage forms and the rectified waveform

Figure 3.8: The emitting and returning current pathways for the rectifier

load as shown in figure 3.7 becomes readily apparent.

However, we have yet to explain why a permanent magnet has been chosen as the

flux source as opposed to an electro-magnet. The reason again concerns the scaling

laws that affect these devices ( Cugat et al. (2003)). For a permanent magnet with

volume v and polarization J , the scalar potential of its magnetic field will be

V pP q v

4πµ0

~J ~rr3

(3.3)

As can be seen then, the vector potential is proportional to length

V pP q 9 L

and therefore scales directly with a scaling factor, k. The magnetic field created by

73

Page 90: The Design of a Micro-turbogenerator

the magnet, ~H, however is equal to the gradient of the scalar potential.

~H ~grad V (3.4)

A gradient is indirectly proportional to length, so the gradient of the potential func-

tion is independent of length and therefore the magnetic field strength is constant.

This shows that as a permanent magnet system is reduced in size by a factor of

k, the same magnetic field will be observed at geometrically similar locations. The

induced emf, ε, however is proportional to dΦdt

, where Φ is proportional to the area

and field strength. But because the magnetic field strength will be constant as just

shown, the flux will be proportional to the scaling factor squared.

dΦ 9 k2

The induced emf is also proportional to the time rate of change of this magnetic

field. As stated previously, the rotational rate of the micro-generator is indirectly

proportional to size in order to maintain constant tip speed. Combining these two

observations results in,

dt9 k

as shown in figure 3.9. Note that this analysis also results in power density remaining

independent of k for a constant current density, because power is proportional to ε2Rand the resistance is indirectly proportional to 1.

P

V9 pdΦ

dtq2RL3

9pkq2

k1

k39 1

For a electro-magnetic current induction machine, we can determine the scaling

effects by looking at the Biot-Savart law and assuming a constant current density.

The Biot-Savart law shows that

~dB µ0I ~dL r

4πr2(3.5)

74

Page 91: The Design of a Micro-turbogenerator

Figure 3.9: Overall effect of scale reduction 1/k on basic magnet interactions forconstant current density, Cugat et al. (2003)

The magnetic field density then scales as,

~B 9 I

L

which results in the magnetic flux scaling as

Φ 9 BA 9 I

LL2 9 IL

For a constant current density

I

A constant I 9 L2

This results in the flux being proportional to the characteristic length cubed.

Φ 9 L3

As before, the rotational rate is indirectly proportional to the scaling factor k. Com-

bining these observations results in

dt9 k3

k9 k2 (3.6)

75

Page 92: The Design of a Micro-turbogenerator

as shown in figure 3.9. The resulting power density is then proportional to k2.

P

V9 pdΦ

dtq2RL3

9pk2q2

k1

k39 k2

For this reason, it is not advisable to use a micro-electro-magnetic generator in con-

junction with micro-turbomachinery, and this is why a permanent magnet generator

was selected for this design.

As stated previously however, the heat dissipation capabilities of electric windings

also scale favorably with k, which explains why permanent magnet generators show

power density improvements as they scale down in size (Arnold et al. (2006a)). The

effect of this added property is shown in figure 3.10.

Figure 3.10: Effect of scale reduction 1/k on magnetic interactions, taking intoaccount increased admissible current density, Cugat et al. (2003)

3.1.2 Magnetic Circuits and Modeling

Now that a permanent magnet generator has been selected, their operation and flux

creating abilities must be described in more detail. As described by Hanselman

(1994), hard magnetic materials exhibit significant hysteresis in their response to

magnetic fields as shown in figure 3.11 and as a result, display a remanence magnet

field strength, Br, without the presence of an external magnet field. By knowing

76

Page 93: The Design of a Micro-turbogenerator

Figure 3.11: A B-H graph for a hard magnetic material

the B-H characteristics of a hard magnetic material, Ampere’s law can be used to

determine a magnets point of operation. Ampere’s law states that

¾~H ~dl µ0i

Applying this law then to the a permanent magnet encased in a ferromagnetic ma-

terial as shown in figure 3.12 , and noting that no current is present in the system,

Figure 3.12: A simple permanent magnet circuit

results in

Hmlm Hklk 0

77

Page 94: The Design of a Micro-turbogenerator

Hence

Hm lklmHk

Because there is a positive magnetic B field in the encasing material, there must also

be a positive H-field as per ~B µ0~H. From Ampere’s law then, the magnetic H

field in the permanent magnet must be negative and the magnet is operating in the

second quadrant of its B-H curve.

Consider a similar magnetic system with an air gap as shown in figure 3.13.

Ampere’s law will result in

Figure 3.13: A simple permanent magnet circuit with an air gap

Hm lklmHk lg

lmHg (3.7)

Assuming that there is no leakage flux, neglecting fringing of the magnetic field at

the air gap, and assuming some representational area for the encasing material Ak

along its length, by Gauss’s law we will have

Φ BmAm BgAg BkAk

Recall the relationship between the magnetic B field and the magnetic H field

~B µrµ0~H

78

Page 95: The Design of a Micro-turbogenerator

The magnetic H fields then take on the following values

Hg BmAmµr,gµ0Ag

Hk BmAmµr,kµ0Ak

Plugging these results into 3.7 gives

Hm lgBmAmµr,gµ0Aglm

lkBmAmµr,kµ0Aglm

(3.8)

The relative permeability of free space of air, µr,g is approximately 1, whereas for high

permeability materials this value, µr,k, can exceed 10,000. Looking at equation 3.8,

we see that the second term can be mostly neglected. Making this simplification

results in

Hm lgBmAmµr,gµ0Aglm

(3.9)

which can be rearranged to give

Bm

µ0Hm

µr,gAglmAmlg

(3.10)

This is the slope of the operating line for magnetic system, which in conjunction with

the magnetic properties of the permanent magnet, will determine its operating point

and flux density emission as shown in figure 3.14. The basic operating principles

of permanent magnets have been presented. With this information, their operation

will now be presented in a slightly different manner that is highly conductive towards

magnetic modeling.

Magnetic circuits are similar to electric circuits and can be modeled using similar

concepts, this basic concept will be demonstrated here. Applying Ampere’s law to

a geometrically symmetrical object such as a coil wound toroid in figure 3.15 we see

that

H l Ni

79

Page 96: The Design of a Micro-turbogenerator

Figure 3.14: The B-H operating conditions for a permanent magnet system

Figure 3.15: A toroidal ferromagnetic core wrapped within a current conductingwire, Nave (2011)

80

Page 97: The Design of a Micro-turbogenerator

Combining this with the relationship between magnetic fields, B µH, where µ µrµ0, results in

B µNi

l

We can multiply this term by the cross sectional area to get

BA Φ µA

lNi (3.11)

Thus there is a linear relationship between some constant parameters of the system,

µAl

, some forcing function, Ni, and the flux, Φ. This relationship is analogous to

Ohm’s relationship between voltage, current, and conductance, and it is very advan-

tageous then to model magnetic circuits in a similar manner. The parameters of the

system, µAl

, will be termed permeance, P, and the forcing function Ni will be referred

to as the magneto motive force, MMF. We require a method then for determining

the MMF of a permanent magnet.

Looking at figure 3.16, a close-up of the B-H curve, and knowing the relation-

Figure 3.16: A close up of the B-H curve for a hard magnetic material

ship between the magnetic H field, the magnetic B field, magnetization M, and the

81

Page 98: The Design of a Micro-turbogenerator

permeabilities,

B µrµ0H M

we can see that the slope is µr as expected. Through geometry then, we can write a

new expression for Bm.

Bm Br µrµ0Hm

If we multiple both sides by the area, the result is.

φm BrAm µrµ0HmAm

Thus the actual flux exiting the magnet is equal to the highest capable flux that

magnet could provide in a zero reluctance environment (φr BrAm) minus some

term that is related to magnetic H field imposed by its environment as per Ampere’s

law. By multiplying the above equation by ll, we get

φm BrAm µrµ0Amlm

Hmlm

Substituting in the values for magnetic permeability and the forcing function from

Ampere’s law, this becomes

φm φr PmF

In an analogous fashion to electric circuit theory then, a permanent magnet can

be modeled as a constant current source in parallel with a resistance as shown in

figure 3.17. Thus the same rules that model current sources in electric circuits can

be used to model permanent magnets in magnetic circuits. As example, looking back

at our simple magnetic circuit with the air gap (fig. 3.13), this could be magnetically

modeled as shown in figure 3.18.

With magnetic circuit knowledge, and an ability to model permanent magnets,

we can now look at a magnetic circuit for a micro-generator. Such a a circuit is

shown in figure 3.19. If we neglect leakage flux, and assume the permeabilities of the

82

Page 99: The Design of a Micro-turbogenerator

Figure 3.17: A circuit model for a permanent magnet

Figure 3.18: A circuit model for a permanent magnet enclosure with an air gap

back iron and stator iron are infinite, this circuit can be simplified and redrawn as

shown in figure 3.20. From this, two magnetic circuit equations can be written

φm Pg2F φr

Pm2 Pg

2

F (3.12)

Equating for the magneto motive force and setting these equations equal to each

other results in

2φmPg

2φr

Pm Pg(3.13)

Assuming that the magnet area and area gap area are the same, solving for the

83

Page 100: The Design of a Micro-turbogenerator

Figure 3.19: A circuit model for a planar permanent magnet generator,Herrault(2009)

system flux results in

φm φrPgPm Pg

φrµ0Alg

µ0Alg

µ0µrAlm

φrlg

1lg µr

lm

φr1

1 µrlglm

(3.14)

This provides a simple equation by which we can approximately estimate the flux

through a pole-pair coil for initial calculations. From this parameter as well as the

Figure 3.20: A simplified circuit model for a planar permanent magnet generator

84

Page 101: The Design of a Micro-turbogenerator

design RPM, we can begin to estimate and determine system performance.

As shown, the observed flux through the coils depends on the remanence flux

of the magnet, which is equal to the flux through the magnet in a zero reluctance

environment. Permanent magnet material such as SmCo is magnetized by placing

the material in a strong magnetic H-field, and then removing this H field, allowing

the material to recoil along its hysteresis path as shown in figure 3.21. The magnetic

Figure 3.21: A B-H curve for a permanent magnet material undergoing magneti-zation

material is typically subjected to strong fields by placing it in patterned permeable

material wrapped in loops of an electric conductor as shown in figure 3.22 and then

running large currents through the coils.

The mechanism by which a permanent magnet generator creates magnetic flux

has been presented as has as a method to quantify the flux leaving the permanent

magnet in a simplified system. Knowing this, and other system parameters such

as RPM, the open-circuit voltage can be determined using Faraday’s law. With

this, many properties of the generator such as maximum power output and electrical

85

Page 102: The Design of a Micro-turbogenerator

Figure 3.22: A magnetizing head used to create permanent magnets, Gilles et al.(2002)

efficiency can be determined.

3.1.3 Electric Circuits and Modeling

A single phase of a permanent magnet generator can be modeled, as shown in fig-

ure 3.23, as an AC voltage source with an internal resistance Rs, corresponding the

resistance of the generator coils in the stator, in series with a load, RL. The inductive

Figure 3.23: A single phase AC circuit

effects can typically be ignored because the stator coils in micro-generators have so

86

Page 103: The Design of a Micro-turbogenerator

few turns, and the inductance of a coil is proportional to N2.

With a fixed voltage source and an internal resistance, the power properties of

the device take on interesting characteristics. The first of these is that there is a

maximum amount of power that can be transferred to a load. By looking again at

figure 3.23, we see that the power transfered to a load is

PL i2RL (3.15)

where

i V

RS RL

(3.16)

Combining these two equations results in

PL V 2RL

pRS RLq2 (3.17)

Taking the derivative of this equation with respect to RL, setting it equal to zero,

and solving for RS gives RL RS. Therefore, the load resistance should be set to

the phase resistance in order to maximize power transfer. Plugging this condition

into equation 3.17 results in

PL,max V 2

4Rs

(3.18)

Therefore, there is a load resistance value above or below which power transfered to

the load will be reduced. This can be easily deduced by looking at extreme cases.

If the load resistance was close to infinite, then no current would be flowing and

therefore there would be no power. If the load resistance was close to zero, then the

current would be set by the phase resistance, and no power would be utilized by the

load.

The efficiency can also be determined from this information. Taking

PG V 2

RS RL

(3.19)

87

Page 104: The Design of a Micro-turbogenerator

and combining this with with equation 3.17, we get

η PLPG

V 2RL

pRS RLq2pRS RLq

V 2 RL

RS RL

1

1 RSRL

(3.20)

Therefore, the higher the load resistance in comparison to the phase resistance, the

higher the efficiency. Because of this, there seems to be a trade off between power

and efficiency, because as the load resistance is increased, the efficiency increases

while power output drops.

This relationship can be quantified. Noting from equation 3.20 that

pRS RLq RL

η(3.21)

and plugging this into equation 3.17, we get

PL V 2RLRLη

2 V 2η2

RL

(3.22)

We wish to have the load power be only a function of the non-varying parameters V

and RS and the efficiency. So using equation 3.21 again, we get

RL ηpRS RLq RL ηRS

1 η(3.23)

Plugging this result into equation 3.22 gives

PL V 2η2 1 η

ηRS

V 2

RS

ηp1 ηq (3.24)

The general graph for this relation takes on the shape shown in figure 3.24, although

the values for the vertical axis will be determined by the system voltage and internal

resistance. As can be seen from the equation, for a given efficiency, higher generator

voltages will increase the power output across the load. Conversely, for a given power

88

Page 105: The Design of a Micro-turbogenerator

Figure 3.24: Load power as a function of efficiency

output, higher voltages will increase electric efficiency. And as with any engineering

system there is a trade off in design. In this case, selecting the load resistance will

be a trade off between power density and energy density as a result of the relation

between power output and efficiency.

3.1.4 Magnetic Materials and Configuration

Thus far, most of the terms in our equations have been variables and they will change

between different systems. However, permanent magnet materials have physical and

magnetic properties that will stay constant in various designs, and these properties

will be discussed here.

As shown from previous sections, a key parameter in the determination of power

output and efficiency is voltage. This in turn is dependent on dφdt

. As shown in

equation 3.14, dφ is dependent on the remanence flux, dφr (which is equal to BrA),

and µr. The maximum remanence flux density, Br, is a magnetic property of ma-

terial and µr is the recoil line for magnetic materials in the second quadrant as

shown in figure 3.16. These values vary between the different permanent magnet

materials such as AlNiCo, NdBFe(Neodymium Boron), SmCo(Samarium Cobalt),

89

Page 106: The Design of a Micro-turbogenerator

etc. Demagnetization curves for these materials are shown in figure 3.25. As can

Figure 3.25: Demagnetization curves for various magnetic materials, Gieras et al.(2004)

be seen from the figure, ferrites are not typically used due to their low remanence.

AlNiCo materials on the other hand have a high remanence, but their value of µr

is extremely high, resulting in low flux values in typical magnetic circuits. Of the

two remaining choices, NdFeB and SmCo, NdFeB has the higher remanence. How-

ever, as shown by Herrault et al. (2008) in figure 3.26, the performance of NdFeB

degrades much more quickly than SmCo in elevated temperature environments, and

thus SmCo magnets are better suited for working in a micro-engine environment.

And due to this temperature dependence, the generator is forced onto the cold side

or the compressor side of micro-engines, and therefore the magnets do not experience

the high temperatures of the combustion gases.

Another important material property for permanent magnet material is the yield

strength. This is a result of the high stress environment due to rotation. Recalling

90

Page 107: The Design of a Micro-turbogenerator

Figure 3.26: A comparison between SmCo and NdFeB generators as a function oftemperature, Herrault et al. (2008)

from centripetal motion that

F mrω2

and noting that rω 9 1 for a turbomachine system, we can see that the forces

experienced by the material scale indirectly proportional to the characteristic length.

F 9 1

r

Therefore, the materials in micro-turbogenerators experience higher forces than their

macro-size counterparts.

Arnold et al. (2005) studied this effect to determine the maximum RPM values

that these devices were capable of sustaining. For an annular magnet the maximum

radial stress and hoop stress are

σr,max 3 ν

8ρω2 rR2 R1s2 (3.25)

σθ,max 1

4ρω2

p3 νqR22 p1 νqR2

1

(3.26)

91

Page 108: The Design of a Micro-turbogenerator

Given that the ultimate tensile strength of SmCo is 35 MPa (Arnold et al. (2005)),

this sets the maximum achievable RPM before failure. For an outer diameter of

approximately 10 mm, the maximum achievable RPM was reported to be 140,000

RPM. In order to increase the maximum allowable RPM, an circumscribing metal

adaptor was placed around the rotor. A similar example can be seen in figure 3.27.

This adaptor can not only pre-strain the magnet, but also provides additional stiff-

Figure 3.27: An annular magnet with a mounting adaptor, Herrault et al. (2008)

ness. With an adaptor, the speed increased to 325,000 RPM without failure, where

the maximum speed was limited by the RPM of their driving turbine, not the stress

levels of the magnetic rotor.

Another configuration option is to use discrete magnetic pieces that are arranged

around the periphery as shown in figure 3.28. There is a maximum stress level

penalty associated with this configuration as there is no ring stiffness in the magnet.

However, this configuration presents a method to mitigate rotor imbalance through

proper selection and placement of the magnetic pieces around the rotor. In the case

of an annular ring, balancing the rotor presents a major challenge as it is difficult to

address imbalances that are present in the material. The main advantage of discrete

pieces however is that they can be much more easily magnetized.

Another key material property in magnetic systems is permeability, µ. As shown

92

Page 109: The Design of a Micro-turbogenerator

Figure 3.28: Discrete magnet pieces arranged peripherally around the generatorrotor, Herrault (2009)

in the magnetic modeling section, through any material

φ MMF

P MMF

µAl

and therefore permeability in magnetic material is equivalent to conductivity in

electrical circuits, and should be maximized in order to increase the magnetic flux

through the system. An additional important parameter is the saturation flux den-

sity. Looking at a typical B-H curve as shown in figure 3.29, the permeability is

constant, that is there is a linear relationship between H and B, below certain values

of magnetic flux. After this point however, the curve begins to shallow, signifying that

the permeability has decreased substantially. This transition point is termed the sat-

uration density. Typical materials used as back irons are NiFe and Hiperco(FeCoV).

NiFe has a saturation flux density of 0.8T (Laughton and Warne (2003)) and Hiperco

a saturation flux density of 2.4T (Arnold et al. (2006b)). However, NiFe is often used

in micro-systems, because techniques for its sputter deposition are mature.

93

Page 110: The Design of a Micro-turbogenerator

Figure 3.29: A B-H curve for a magnetic material

3.2 Device Performance

3.2.1 Device Geometry and Parameters

Now that the concepts of permanent magnet generators have been presented, a micro-

generator will be designed so that its performance can be estimated. The engine will

have a general arrangement as shown in figure 3.30. The advantage of this design is

Figure 3.30: The arrangement of an uncoupled micro-turbogenerator, Pelekieset al. (2010)

94

Page 111: The Design of a Micro-turbogenerator

that the generator is on a different shaft than the core turbomachinery. This allows

the turbomachinery to operate at its high design RPM, while allowing the generator

to operate at a lower RPM within the material limits of its permanent magnets.

Following the geometry of Herrault et al. (2010), as shown in figure 3.27, the

generator magnets will be designed to have an outer radius of 5 mm and an inner

radius of 2.5 mm. Experimental evidence showed an RPM of 200,000 can be reached

before mechanical failure. Thus, for a factor of safety, a design RPM of 175,000 is

chosen. A micro-turbine was designed to give 10W of shaft power at this speed. The

magnet thickness will be 0.5 mm, the air gap between the magnets and the coils will

be 100 µm thick, and the coils will be be 200 µm thick. The thickness of the back

iron will be determined at a later time in order to avoid saturation. This design will

also have a different key feature than previous designs. As opposed to having one

coil layer, multiple layers will be stacked in series as shown in figure 3.31 in order to

increase system voltage and efficiency.

The effect of this change can easily be seen intuitively by looking at its imple-

mentation in figure 3.32. For every increased layer (n), the output voltage of the

generator will increase at an exponentially decaying rate, and the resistance will in-

crease linearly. Therefore at first, and for a constant generator power output, when

the incremental voltage increase is high per additional layer, this higher voltage ad-

vantage will outweigh the increased resistance. For very high layer counts however,

additional coil layers will be far away from the magnets, and therefore will not in-

crease voltage significantly. The resistance however will still be increasing linearly,

and therefore the performance will be hampered by further increasing the coil layer

count. This was numerically simulated in Matlab for this generator design and the

results can be seen in figure 3.33. The efficiency was shown to begin decreasing at a

layer count of approximately 30. Regardless, the performance gains were not signifi-

cant beyond 5-10 layers and cannot justify the added system complexity. In addition,

95

Page 112: The Design of a Micro-turbogenerator

Figure 3.31: A permanent magnet generator with three coil layers, modifiedfrom Herrault (2009)

Figure 3.32: A representational circuit diagram for a generator with a variablelayer count, n

96

Page 113: The Design of a Micro-turbogenerator

Figure 3.33: Effect of layer count on system parameters

this simple model is not taking into account losses from other mechanisms such as

eddy currents that would further erode performance with higher layers counts. The

thickness of the inter-conducting gap between the coil layers was set to 40µm.

Another change for this design involves the removal of the stator back iron. This

is done in order to avoid significant eddy current and hysteresis losses, at the expense

of less output voltage and therefore power density. An empirical analysis done with

the generator designed and built by Yen et al. (2008a), shows that the effective path

length for the flux in the stator is 736 µm.

The remanence flux density for the magnets was assumed to be 1T as a result of

slightly elevated temperatures. Following the optimization analysis done by Arnold

et al. (2006a), the magnets were patterned with 4 pole pairs, P, and the windings

in the generator have 3 three turns per pole, N. As stated earlier, the generator is

designed with three electrical phases, F. The coils of a single phase can be seen in

figure 3.34.

97

Page 114: The Design of a Micro-turbogenerator

Figure 3.34: A single phase of the 4 pole-pair, 3 turns per pole stator, Herrault(2009)

The widths of the conductors must be approximately estimated so that the phase

resistance can be determined. From geometry, the mean radial width can be deter-

mined from

wr,mean 2πRmean

p2NqFP

which results in a value of 300µm. The widths of the outer and inner segments will

be set to 40% of the mean radial width. The thickness of the radial segments is by

definition equal to the specified thickness of the coils. The inner and outer segments

however must be equal to half of this value in order to accomomdate physical cross-

overs from the different phases. The length of the radial segments is set equal to

the radial length of the magnets, 250 µm. Lastly, the length of the inner and outer

segments can again be determined from geometry

li,o 2πri,o2P

and are equal to 196µm and 393µm respectively. For the power electronics, the use

of active MOSFETs will be assumed with an effective resistance of 0.10 ohms.

With a defined geometry and set system parameters, the performance of the

device can be estimated.

98

Page 115: The Design of a Micro-turbogenerator

3.2.2 Flux and Induced Voltage

In order to more accurately determine system voltage and the generated flux through

the magnets, analytical results from Das et al. (2006) will be used. Das analytically

modeled three phase permanent magnet micro-generators that were very similar to

our design. His model worked by solving 2D Maxwell’s equations through planar

continuum layers as a function of radius (figure 3.35). His results compared very

Figure 3.35: The planar continuum layers used in Das’s analysis, Das et al. (2006)

well with 3D FEA simulations.

Assuming a square wave magnetization profile (cm 0 in figure 3.36), infinite

permeabilities for the back iron and stator iron, and B and H fields independent of

radius, Das’s equation for output voltage reduced to

V0 R2o R2

i

TalTal Tcl Tag

Br

NPω (3.27)

This equation was slightly modified for our analysis because the stator back iron

was removed and an inter-conductor gap was introduced between coil layers. In the

99

Page 116: The Design of a Micro-turbogenerator

Figure 3.36: A magnetization profile for an annular permanent magnet, Das et al.(2006)

denominator, a term for the length of the flux path in the tangential direction was

added, Tfp, as was a term for the inter-conductor gap, Tg. The length of the average

flux path at the mean radius can be determined by geometry.

c1

2 1

2

1

2p2πRo 2πRiqp2P q

(3.28)

An additional 12 is present to be consistent with Das’s analysis, where only half of

the magnetic circuit is analyzed (the other half is symmetrical). For the value of c,

an empirical analysis was done with a similar generator constructed by Yen et al.

(2008a) which also lacked a stator back iron. The value of c came out to 1.5.

The equation for voltage was also modified by another constant, k, as a result

of non-uniform trapezoidal magnetization. Comparing the value predicted by using

this model with the experimentally tested generator by Arnold et al. (2006b) resulted

in a value of 0.72 for k.

The resulting single phase voltage values are shown in table 3.1. Note that this

flux was determined assuming an infinite permeability in the rotor back iron. In order

for this assumption to be approximately true, the back iron must not be saturated.

This is the criteria which sets the back iron thickness. Looking at figure 3.37, and

100

Page 117: The Design of a Micro-turbogenerator

Table 3.1: Voltage results

n V°V

°Vwye

1 0.77 0.77 1.33

2 0.68 1.45 2.50

3 0.61 2.06 3.56

4 0.56 2.61 4.52

5 0.51 3.13 5.41

6 0.47 3.59 6.22

7 0.44 4.03 6.98

8 0.41 4.43 7.68

9 0.38 4.82 8.34

10 0.36 5.17 8.96

assuming that Ro ¡¡ Ri such that the tangential flux density is independent of

radius we can determine the following relations. The area of the flux leaving the

permanent magnet is

Am 1

2

πpR2o R2

i q2P

(3.29)

Note that only half of the pole area is used, because each half sees flux traveling in

Figure 3.37: A schematic for approximating the back iron thickness

101

Page 118: The Design of a Micro-turbogenerator

a different tangential direction. Therefore, the total flux entering the back iron is

φ AmBm (3.30)

The area of the back iron normal to the tangential flux is

Abi pRo Riqtbi (3.31)

where tbi is the thickness of the back iron. The flux density through the back iron

with our approximations is therefore

Bbi φ

Abi(3.32)

Combing all of these equations results in

Bbi φ

tbipRo Riq (3.33)

The maximum flux density that the back iron should receive is its saturation flux

density, the value of Bbi therefore should not exceed this. Plugging in this saturation

value into the above equation and re-arranging for the back iron thickness gives an

approximation for how thick the back iron should be to avoid saturation.

tbi φ

BsatpRo Riq (3.34)

Assuming a Hiperco50 (FeCoV) back iron which has a flux saturation density of

2.4T, the thickness of our back iron becomes 161µm.

3.2.3 Generator Coil Resistance

In order to determine power outputs and efficiency, a knowledge of the coil resistance

is required. Looking at figure 3.34 we can determine from geometry the areas and

path lengths required to determine resistance.

102

Page 119: The Design of a Micro-turbogenerator

From the figure, we can see that there are 2 radial conductors per pole pair, 1

outer conductor per pole pair, and 1 inner conductor per pole pair. Recalling that

the equation for resistance is,

R ρl

A(3.35)

and assuming copper conductors (ρ 1.68 108Ohm m), we can determine the

resistance of each radial, inner, and outer segment, and then multiply this by the

number of segments for each type and sum the values to get the resistance per phase.

Doing this results in the following values for resistance per phase per layer.

Rr 15.4 mΩ

Ri 30.2 mΩ

Ro 60.5 mΩ

Rtotal 106 mΩ

Two additional adjustments must be made. The first is to multiple the coil resistance

by a correction factor, k, to account for the difference between the model predictions

and reported values. These differences stem largely from manufacturing defects, vias,

and other geometries that are difficult to model. Applying this same technique to the

generator reported by Herrault et al. (2010), resulted in a correction factor k of 3.

With this adjustment, the phase-phase winding resistance for a single layer becomes

Rphph 0.64 Ω

The second adjustment is to add an effective resistance as a result of the MOSFET’s

in the power electronics. This was assumed to be 100 mΩ.

103

Page 120: The Design of a Micro-turbogenerator

3.2.4 Maximum Power Transfer Capabilities

With knowledge of the resistance and voltage values, the maximum power dissipation

abilities of the generator can be be determined as a function of layer count. This

information has two practical importances. The first is that, as shown in figure 3.24,

the ratio between maximum power to actual load power is strongly related to electric

efficiency. The other reason is that it is necessary to determine the minimum number

of generator layers for the generator to operate at its design RPM. As example,

assuming only electrical losses, for a fixed value of voltage and coil resistance, there

will be a maximum amount of power that the generator can dissipate (by setting

RL = 0). If this amount is less than the input shaft power, then by definition the

torque being applied to the generator from the turbine will be greater than the torque

applied to the generator from the electrical current. As a result, the generator will

accelerate above its design RPM and potentially fail. Thus, there is a minimum layer

count required such that it is within the ability of the generator to utilize or dissipate

the input shaft power and maintain its design RPM.

Recall that for this generator, the phase to phase resistance and power electronics

resistance is

Rphph 0.64Ω Rpe 0.1Ω

Therefore, the total phase-phase resistance will be

Rt nRphph Rpe (3.36)

The maximum power that can be delivered per phase is observed when the load

resistance is set to 0. Therefore, assuming a wye connected 3-phase generator system,

the maximum power that the generator can electrically dissipate is

Pmax 3V 2

wye

Rt

(3.37)

104

Page 121: The Design of a Micro-turbogenerator

Table 3.2: Maximum power dissipation

n Vwye pV q Rt pΩq Pmax pW q1 1.33 0.74 7.16

2 2.50 1.37 13.69

3 3.56 2.01 18.94

4 4.52 2.65 23.21

5 5.41 3.28 26.71

6 6.22 3.92 29.61

7 6.98 4.56 32.03

8 7.68 5.19 34.07

9 8.34 5.83 35.79

10 8.96 6.47 37.25

The results as a function of layer count can be seen in table 3.2.

As can be seen, with only a single coil layer, the generator does not have the

ability to dissipate 10W of shaft power. Thus, ignoring non-electrical losses for the

time being, more than 1 coil layer would be required to maintain the generator at

its design RPM.

3.2.5 Loss Mechanisms

Eddy Currents and Hysteresis Losses

In addition to the electrical losses in the resistive elements, there exist other losses

associated with the magnetic character of the generator. One of these are hysteresis

losses. As shown in figure 3.11, magnetic material will follow different paths as it

continually observes a changing H-field. The area encompassed by the hysteresis

curve is proportional to the energy loss, and therefore power loss is proportional to

this area times the frequency. However for our generator, we have decided not to use

a soft-magnetic material as our back iron, and therefore these effects are essentially

zero in the stator. The cost of this decision is higher magnetic reluctance, reduced

flux, lower voltage, and therefore lower electrical efficiency for the same load power.

105

Page 122: The Design of a Micro-turbogenerator

As for the rotor, the magnetic material sees a non-time varying flux, and therefore

does not experience hysteresis.

The other type of magnetic losses present are eddy current losses. The mechanism

behind eddy currents are the same as those responsible for creating a voltage in the

generator coils. Due to Faraday’s law, when a small area in a material is experiencing

a changing magnetic field, voltages will be induced around the material. For non-

conducting material, current cannot flow so no power will be dissipated, but in the

case of conducting material, currents will be present and dissipate power. These

losses can be reduced through the use of laminations as shown in figure 3.38, which

are non-conducting materials that act as barriers to current flow. Again however, we

Figure 3.38: Laminations embedded within a magnetic conducting stator

have chosen to use a non-magnetic non-conducting stator core, and therefore these

losses are avoided. However, these losses are present within the radial segments of the

conductors themselves as shown in figure 3.39 Again, these losses can be minimized

by introducing laminations within the radial conductor segments. An analytical

analyses by Das et al. (2006) showed that these losses can be approximated as

Pprox p2FNClamqσc96TclpRo Riq

πRm

3NClam

3

ω2B2 (3.38)

where Clam refers to the number of laminations in the copper conductors. Our design

has only one lamination (zero non-conducting barriers). The total eddy current losses

within the conductors as a function of layer count are shown in table 3.3.

106

Page 123: The Design of a Micro-turbogenerator

Figure 3.39: Eddy currents within a radial conducting segment with and withoutlaminations, Das et al. (2006)

Table 3.3: Eddy current losses

n Pproxpnq pW q °Pprox pW q

1 0.14 0.14

2 0.12 0.26

3 0.092 0.35

4 0.076 0.43

5 0.064 0.49

6 0.054 0.55

7 0.047 0.59

8 0.041 0.63

9 0.036 0.67

10 0.032 0.70

107

Page 124: The Design of a Micro-turbogenerator

Ohmic Losses and Power Matching

The remaining loss mechanism associated with electromagnetic properties of the

generator is ohmic losses as a result of current running through the coils of the

generator stator and the power electronics. This loss is simply the result of Ohm’s

law and is expressed as

Ps,p.e. Fi2phphRt (3.39)

where F again refers to the number of phases. However, the current through the

coils depends on the load resistance, which has yet to be determined.

In order to determine the load resistance, a power balance is required. That is,

the shaft power input into the generator must equal the load power plus all of power

losses due to the electromagnetic loss mechanisms.

Pshaft Pload Pprox Ps,p.e. (3.40)

This is absolutely essential because it ensures that the magnetic rotor operates at the

design RPM and neither accelerates or decelerates. Every variable in equation 3.40

has been defined with the exception of the load power. Therefore, the resistance of the

load must be set such that this equation holds true. We can re-arrange equation 3.40

and define a new variable, Pcircuit.

Pshaft Pprox Pload Ps,p.e. Pcircuit (3.41)

Pcircuit then represents the amount of power that must be dissipated electric circuit

components of the system as shown in figure 3.40, whether that be in the stator

windings or the load, in order to balance the system. Therefore the required load

resistance can be determined from

Pcircuit FV 2

wye

RL Rs,p.e.

108

Page 125: The Design of a Micro-turbogenerator

Figure 3.40: A representational circuit for power balance

RL FV 2

wye

PcircuitRs,p.e.

and the required current to balance the system will become

Iphph VwyeRs,p.e. Rload

With this, we can now solve equation 3.39 for power dissipation in the coil windings.

The results are shown as a function of layer count in table 3.4. Note that the

Table 3.4: Ohmic losses in the stator windings and power electronics

n RLpnq pΩq ipnq pAq Ps,p.e.pnq pW q1 -0.20 2.47 13.51

2 0.56 1.29 6.90

3 1.95 0.90 4.89

4 3.78 0.70 3.93

5 5.96 0.58 3.37

6 8.38 0.51 3.0

7 10.99 0.45 2.75

8 13.74 0.41 2.56

9 16.59 0.37 2.42

10 19.50 0.35 2.31

resistance value for a single layer is negative, or alternatively that the ohmic losses

109

Page 126: The Design of a Micro-turbogenerator

in the stator and power electronics are greater than the shaft power. This indicates

that generator system has no means, with only a single coil layer, with which to

dissipate or utilize sufficient shaft power to maintain its RPM.

3.2.6 Power Output, Efficiency, and Layer Count

With all of the loss mechanisms having been determined, we can now look at load

power and system efficiency as a function of layer count. With this information,

we can determine which generator design should be physically constructed based on

trade-offs between system performance, complexity, and cost as a result of increased

layer counts.

With generator efficiency defined as,

ηgen PLPshaft

(3.42)

the results are as shown in table 3.5. As before, notice that this generator cannot

Table 3.5: Load power and efficiency as a function of layer count

n PL pW q ηgen %

1 - -

2 2.82 28

3 4.73 47

4 5.62 56

5 6.12 61

6 6.43 64

7 6.63 66

8 6.78 68

9 6.89 69

10 6.97 70

function at this RPM with only a single coil layer.

In addition, as expected from figure 3.33, there are diminishing returns to in-

creasing the layer count. As such, a generator with 6 layers, a power output of 6.4

110

Page 127: The Design of a Micro-turbogenerator

W, and an efficiency of 64% seems to be the best choice with regard to the system

trade offs.

This generator was designed within the physical parameter limits of micro-generators

in the literature. The performance can be improved by venturing outside of this de-

sign space (which may require fabrication innovations). As an example, assume

superior coil fabrication reduced the coil correction factor from 3 to 2, the magnet

thickness was increased from 0.5 mm to 1mm, the magnetization profile was improved

such that its correction factor went from 0.72 to 0.95, 3 laminations were placed in

the radial segments, and non-conductive low-hysteresis high-frequency ferrites were

used as the stator back iron in order to reduce the magnetic reluctance of the flux

path. With such a generator, the optimal efficiency would be observed with only two

coil layers and the generator would have an efficiency of 93%.

3.3 Conclusion

The concepts that govern electric generators were presented. Scaling laws were also

shown that demonstrated why permanent magnet generators offer superior perfor-

mance at small size when compared to electromagnetic induction machines.

The principles of permanent magnet performance were explained as were the

concepts of magnetic circuits. A 1 cm diameter permanent magnet generator was

designed and the performance values were determined. The device demonstrated an

output power of 6.4 W with an efficiency of 64% and a coil layer count of 6. With

specific improvements, the device demonstrated 9.3 W at an effeciency of 93% and

a coil layer count of only 2.

The performance of micro-turbogenerators using these two generators can be

compared against batteries in terms of energy density. Mattingly et al. (2006) gives

111

Page 128: The Design of a Micro-turbogenerator

the performance of an ideally recuperated Brayton cycle as

ηtherm 1 PRk1k

T4T2

Assuming a compressor efficiency of 50%, a core turbine efficiency of 50%, and a

power turbine efficiency of 50% (all obtained using high speed radial turbomachin-

ery), a pressure ratio of 1.5, and a turbine inlet temperature of 1000K, results in

an thermal efficiency of 8.28%. As shown in figure 3.41, with generator efficiencies

of 64% and 93% from our low-end and high-end designs, the energy density of these

engines would be 3x and 4.5x greater respectively than lithium-ion batteries

Figure 3.41: Energy density of our designed micro-turbogenerators relative to Li-ion batteries

112

Page 129: The Design of a Micro-turbogenerator

4

Experimental Results of an Ejector DrivenMicro-turbogenerator

4.1 Introduction

This section demonstrates an alternative thermodynamic cycle that was used to

overcome the challenges of the Brayton cycle at the micro scale. In this cycle, the

engine is designed around static pumping devices, an injector and an ejector [Gardner

et al. (2010a)]. Both are based on the same fundamental principle: a high momentum

motive fluid is mixed with a low momentum suction fluid, resulting in a discharge

fluid with less overall momentum and thus a higher pressure. The ejector is used

primarily to create a pressure gradient across the turbine, while the injector pumps

liquid into a high pressure boiler. The benefit of these components lies in their static

and turbo-machine independent operation. Unlike a turbo-compressor, the ejector is

uncoupled from the turbine and thus provides a pressure gradient across the turbine

regardless of the turbine’s rotor speed so long as heat is applied to the boiler.

This eliminates many problems associated with the startup phase and guarantees

that the cycle will close at any operating efficiency. In addition, because there are no

113

Page 130: The Design of a Micro-turbogenerator

Figure 4.1: Control volume of the ejector mixing region

moving parts associated with these components, they can be more easily fabricated

as their manufacturing tolerances are not as critical as those associated with rotating

micro-turbomachinery components. Lastly, the suction provided by the ejector can

be utilized to prime the hydrostatic gas bearings, removing the need for an external

pressure source.

This section will discuss both the thermodynamic model concerning this type

of cycle and the preliminary experimental results of an ejector-driven micro-turbo-

generator.

4.2 Thermodynamic Cycle

The proposed cycle is derived from a steam locomotive cycle and an after-burning

Brayton cycle (Fig. 2). Combustion takes place downstream of the turbine, and

the generated heat is split between preheating the turbine inlet air with the use of

a recuperator and vaporizing ethanol in the boiler. Note that for this cycle ethanol

114

Page 131: The Design of a Micro-turbogenerator

Figure 4.2: Schematic of an after-burning thermodynamic cycle driven by an in-jector

is initially used both as the fuel and as the motive vapor for both the injector and

ejector. The power from the turbine can be estimated from the incompressible flow

assumption, the ideal gas law, and by recalling that the total pressure at the turbine

inlet is roughly equal to ambient pressure.

9Wt

9ms

ηt∆Ptρ

ηtPamb Pt,s

PambRairTt4 (4.1)

The pressure difference between the ambient and total suction pressures can be

determined by the conservation of mass and momentum for the control volume shown

in Fig. 1, and assuming isentropic expansion in the diffuser such that the exit

dynamic head is approximately zero. The result is

Pamb Pt,s12ρu2

m

σ

2 σ α2σ3

p1 σq2 2ασ2

p1 σq 2ασ

p1 σq

(4.2)

115

Page 132: The Design of a Micro-turbogenerator

where σ is the area ratio and α is the ejector entrainment ratio.

σ AmAd

α 9ms

9mm

The system rejects heat in the ejector discharge flow and by the cooling and conden-

sation of the ethanol vapor. However, the heat rejected in condensation typically far

exceeds the heat rejected from the cooling of the exhaust gases. The overall thermal

cycle efficiency can be approximated as

η 9Wt

9Wt 9mmhfg(4.3)

In addition, by further assuming the turbine power is significantly less than the

rejected heat we can re-approximate the cycle efficiency as

η 9Wt

9mmhfg(4.4)

Combining equations (4.1) - (4.4) we obtain

η ηtRairTt4ασ

2 σ α2σ3

p1σq2 2ασ2

p1σq 2ασ

p1σq

12ρu2

m

hfgPamb(4.5)

where σ, um, and Tt4 are design parameters (um is a function of boiler pressure).

Picking an a reasonable area ratio of σ = 0.5, the highest cycle efficiencies will

occur for an entrainment ratio of 0.53. Using these values and assuming a turbine

efficiency of 50%, we can approximate the overall efficiencies for different parameters

as shown in Table 4.1.

4.3 Experiment and Results

An experiment was conducted to demonstrate that an ejector powered by ethanol

vapor could create a pressure gradient across a turbine and drive its attached micro-

116

Page 133: The Design of a Micro-turbogenerator

Table 4.1: Efficiency approximations

Case 1 Case 2

PboilerPs

3 30

Tt4 500 K 1600 K

η 1.13% 5.36%

generator to deliver electrical power. The turbine was originally designed and micro-

machined as a radial flow impulse turbine with an NGV outer-diameter of 10 mm

and blade heights of 250 µm (Fig. 4.3). However, due to the configuration and choice

Figure 4.3: The original micro-turbine design bonded to the rotor of a permanentmagnet generator with protruding leads

of materials, we believe eddy current losses in the surrounding material prevented

the turbine from operating on design. A new turbine with a rotor outer-diameter

of 11 mm and blade heights of 750 µm was 3D printed out of ABS plastic for the

final experiments. This allowed the turbine to reach a speed closer to its design

RPM. The ejector was micro-machined with a throat diameter of 719 µm, an area

ratio of 1 : 8, and was driven by ethanol vapor from a conventionally-sized boiler.

The turbine rotor was bonded to a 3-phase Faulhaber 1202-H-006-BH permanent

117

Page 134: The Design of a Micro-turbogenerator

magnet DC motor with the outputs being rectified to DC with the use of 1N5817

Schottky diodes (Fig. 4.4 ). The rectified DC source was then connected across a

Figure 4.4: 3D printed turbo-generator connected to power electronics (boiler andejector not shown)

variable resistor that was adjusted until the maximum power output was obtained.

The conditions at the optimal operating point are shown in Table 4.2. Power from

Table 4.2: Experimental results

Property Units Value

Rotor Speed RPM 27,360

Turbine Pressure Ratio - 1.05

Boiler Pressure atm 15

VDC V 1.49

Turbine Inlet Temp. K 293

Power mW 7.5

the engine was also used to light a row of LEDs (Fig. 4.5)

118

Page 135: The Design of a Micro-turbogenerator

Figure 4.5: LEDs powered by the micro-turbine generator

Note that maximum power does not take place by setting the load resistance

to the equivalent stator resistance as prescribed in the maximum power transfer

theorem. The reason for this is that the maximum power transfer theorem assumes

a fixed voltage source, where as for a turbo-generator the voltage is coupled to the

rotor speed. Therefore, the load resistance must be set such that the turbine can

reach its design speed (Fig. 4.6).

4.4 Conclusion

This section has demonstrated that, with no moving parts, an ejector can produce

a pressure gradient to drive a micro-turbine and generate power. An advantage of

this operating mode is that, unlike a standard Brayton cycle, there is no minimum

required efficiency for the cycle to close and the engine to function. In addition, the

ejector provides a means of creating the pressure gradient required by hydrostatic

gas journal bearings. This will allow the bearings to operate hydrostatically at low

speeds until the RPM increases, thereby allowing hydrodynamic bearing operation.

The manufacturing tolerances of these static pumping devices can also be much less

stringent than those of micro-turbomachinery.

119

Page 136: The Design of a Micro-turbogenerator

Figure 4.6: Turbine shaft power and load power as a function of rotor speed

The thermodynamic cycle was analyzed with both compressible and incompress-

ible flow assumptions, and a basic method of estimating the cycle thermal efficiency

was presented. Experiments were conducted to demonstrate the viability of a power

cycle designed around an ejector-driven micro-turbine.

120

Page 137: The Design of a Micro-turbogenerator

5

Conclusion and Future Work

5.1 Summary and Conclusions

The basic operating principles of micro heat engines were presented. They were

shown to represent a good balance between power density and energy density, and

this characteristic, in comparison to fuel cells and batteries, was shown to make them

good candidates for mobile applications that require both decent power densities and

energy densities.

The basic scaling laws that govern both turbomachinery and permanent magnet

generator power density were presented. It was shown that for turbomachinery,

the power density scales indirectly proportional with the characteristic length of the

system. For this reason, their power densities are observed to increase as they are

reduced in size. For permanent magnet generators, their power density was at first

determined to be scale independent. However, the heat dissipation capability of

the generator windings was shown to increase with reduced size as a result of the

increased ratio of surface area to heat generation. This should allow for more current

density and therefore power density at small sizes at the cost of efficiency or allow for

121

Page 138: The Design of a Micro-turbogenerator

the replacement of cooling elements with power producing elements at small sizes,

thereby increasing power density with a constant efficiency.

Multiple challenges that affect micro-turbogenerators were presented. Of prime

importance, were the efficiency of micro-turbomachinery and the power transfer ca-

pabilities of the generator. The efficiency of micro-turbomachinery was shown to be

important not only for reasons related to energy and power density, but also because

for a Brayton cycle, component efficiencies must meet a specific threshold in order

for the cycle to close and create net power. The power transfer capabilities of the

generator were determined to be important because for a generator converting its

input shaft power to electric power, if the power transfer capabilities do not far ex-

ceed the input shaft power, either the conversion efficiency will be low or the device

will accelerate beyond its design RPM and potentially fail.

The basic operating principles of turbomachinery were developed, with the dif-

ferent components of energy transfer being specifically delineated. The ratio of these

energy transfer mechanisms was shown to relate to turbine reaction and device ef-

ficiency. Loss models were developed to quantify entropy creation from tip leakage,

trailing edge mixing, and viscous boundary layers over the surface of the blades.

The total entropy creation was then related to lost work and turbine efficiency. The

results showed the efficiency and power density of various turbine configurations over

a large range of sizes. For the configurations analyzed, the high speed single stage

reaction turbine showed the best performance. The power density was also shown

to scale linearly as expected over a large range of diameters. However, at very small

scales, the effects of viscous losses superseded the benefits from the scaling laws,

resulting in a peak power density. The single stage reaction turbine was shown to

possess the highest peak power density. The practice of multi-staging was shown to

not be as beneficial at small scales as it is at large scales, because the gains associ-

ated with increased kinetic energy absorption are largely offset by the combination

122

Page 139: The Design of a Micro-turbogenerator

of high viscous losses associated with small scale turbomachinery and the increased

wetted area. The conclusion was that for micro-turbomachinery, high speed, high

reaction, single stage radial designs are most effective due to favorable pressure gra-

dients, low wetted area, and large portion of energy transfer taking place through

lossless centrifugal pressure fields.

The operating principles of generators and power electronics were then presented.

The scaling laws for both permanent magnet generators and electro-magnetic induc-

tion machines were developed and showed that permanent magnet generators should

scale down more favorably than electro-magnetic machines. The basic concepts of

permanent magnet operation were explained, and this was tied together with mag-

netic circuit theory to show the flux generating capabilities of permanent magnets

in planar micro-generators. The flux generation of the magnets was related to volt-

age creation through Faraday’s law, which allowed us to model the generator as

an alternating current source with a fixed internal resistance in an electric circuit.

An analysis was done to show the relationship between efficiency, generator voltage,

internal resistance, and load power.

Models were presented for planar micro-generators to determine output voltage,

internal resistance, electrical losses, and electromagnetic losses based on geometry

and key design parameters. A 3 phase multi-layer permanent magnet generator

operating at 175,000 RPM with an outer diameter of 1 cm was designed and an

efficiency of 65% was shown. A similar device was designed with improved features

that would require fabrication innovations and demonstrated an efficiency of 93%.

Lastly, an ejector driven turbogenerator was designed, built, and tested. A basic

thermodynamic cycle was presented in order to estimate system efficiency as a func-

tion of design parameters. Experiments were conducted showing a power output of

7.5 mW at 27,360 RPM.

123

Page 140: The Design of a Micro-turbogenerator

5.2 Future Work

For future work, geometrically similar turbines should be tested in thermodynami-

cally identical operating conditions to determine the validity of the models presented

here. Alternatively, high fidelity CFD computational studies could be conducted. A

similar loss study model should be done for micro-compressors taking into account

different behavior and flow separation as a result of their unfavorable pressure gra-

dients. The basic loss models however should be similar.

In addition, a multi-layer generator with the dimensions and parameters given

here should be fabricated and tested in order to validate or repudiate the performance

and loss models. Alternatively, 3D FEA studies could be conducted.

After these tasks are completed, an externally supported turbogenerator system

with an integrated compressor, combustion chamber, turbine, generator, and gas

bearing system should be designed, fabricated, and tested. With the lessons learned

from this process, a stand-alone self sustaining device that would not require external

support for bearings, fuel injection, etc., should be designed, fabricated, and run as

the worlds first fully functional micro-heat engine.

124

Page 141: The Design of a Micro-turbogenerator

Bibliography

Arnold, D., Joung, Y., Zana, I., Park, J., Das, S., Lang, J., Veazie, D., and Allen,M. (2005), “High-speed characterization and mechanical modeling of microscale,axial-flux, permanent-magnet generators,” in Solid-State Sensors, Actuators andMicrosystems, 2005. Digest of Technical Papers. TRANSDUCERS’05. The 13thInternational Conference on.

Arnold, D., Herrault, F., Zana, I., Galle, P., Park, J., Das, S., Lang, J., and Allen, M.(2006a), “Design optimization of an 8W, microscale, axial-flux, permanent magnetgenerator,” Journal of Micromechanics and Microengineering, 16.

Arnold, D., Das, S., Park, J., Zana, I., Lang, J., and Allen, M. (2006b), “Micro-fabricated high-speed axial-flux multiwatt permanent-magnet Generators-part II:design, fabrication, and testing,” Journal of Microelectromechanical Systems, 15,1351.

Camacho, A., Gardner, W., Wang, I., Shen, H., Jaworski, J., Gilmore, C., Pelekies,S., Dunnmon, J., and Protz, J. (2010), “Experimental results of an ejector drivemicro turbine generator,” in The 10th International Workshop on Micro and Nan-otechnology for Power Generation and Energy Conversion Applications, PowerMEMS ’10.

Campbell, P. (1996), Permanent magnet materials and their application, CambridgeUniv Pr.

Cebeci, T. and Bradshaw, P. (1977), Momentum Transfer in Boundary Layers,McGraw-Hill Book Company, Inc., New York.

Cebeci, T. and Carr, L. W. (1978), “A Computer Program for Calculating Laminarand Turbulent Boundary Layers for Two-Dimensional Time-Dependent Flows,”Tech. Rep. TM-78470, NASA.

Cugat, O., Delamare, J., and Reyne, G. (2003), “Magnetic micro-actuators andsystems (MAGMAS),” Magnetics, IEEE Transactions on, 39, 3607–3612.

Cumpsty, N. (1989), Compressor aerodynamics, Halsted Press.

125

Page 142: The Design of a Micro-turbogenerator

Das, S., Arnold, D., Zana, I., Park, J., Allen, M., and Lang, J. (2006), “Micro-fabricated high-speed axial-flux multiwatt permanent-magnet generatorsPart I:Modeling,” Microelectromechanical Systems, Journal of, 15, 1330–1350.

Denton, J. D. (1993), “Loss Mechanisms in Turbomachines,” Journal of Turboma-chinery, 115, 621–656.

Epstein, A. H., Senturia, S. D., Al-Midani, O., Anathasuresh, G., Ayon, A., Breuer,K., Chen, K. S., Ehrich, F. E., Esteve, E., Frechette, L., Gauba, G., Ghodssi, R.,Groshenry, C., Jacobson, S., Kerrebrock, J. L., Lang, J. H., Lin, C. C., London,A., Lopata, J., Mehra, A., Mur Miranda, J. O., Nagle, S., Orr, D. J., Piekos,E., Schmidt, M. A., Shirley, G., Spearing, S. M., Tan, C. S., Tzeng, T. S., andWaitz, I. A. (1997), “Micro-Heat Engines, Gas Turbines, and Rocket Engines -The MIT Microengine Project,” in 28th AIAA Fluid Dynamics Conference, PaperAIAA-97-1773.

Frechette, L., Jacobson, S., Breuer, K., Ehrich, F., Ghodssi, R., Khanna, R., Wong,C., Zhang, X., Schmidt, M., and Epstein, A. (2005), “High-speed microfabricatedsilicon turbomachinery and fluid film bearings,” Microelectromechanical Systems,Journal of, 14.

Frechette, L. G., Lee, C., and Arslan, S. (2004), “Development of a MEMS-BasedRankine Cycle Steam Turbine for Power Generation: Project Status,” in 4th In-ternational Workshop on Micro and Nano Technology for Power Generation andEnergy Conversion Applications, Kyoto, Japan, pp. 92–95, Power MEMS ’04.

Gardner, W., Wang, I., Jaworski, J., Brikner, N., and Protz, J. (2010a), “Exper-imental investigation and modeling of scale effects in jet ejectors,” Journal ofMicromechanics and Microengineering, 20, 085027.

Gardner, W., Jaworski, J., Camacho, A., and Protz, J. (2010b), “Experimentalresults for a microscale ethanol vapor jet ejector,” Journal of Micromechanics andMicroengineering, 20, 045019.

Gieras, J., Wang, R., and Kamper, M. (2004), Axial flux permanent magnet brushlessmachines, Kluwer Academic Publishers.

Gilles, P., Delamare, J., and Cugat, O. (2002), “Rotor for a brushless micromotor,”Journal of Magnetism and Magnetic Materials, 242, 1186–1189.

Greitzer, E. M., Tan, C. S., and Graf, M. B. (2004), Internal Flows: Concepts andApplications, Cambridge University Press, Cambridge, UK.

Hanselman, D. (1994), Brushless permanent-magnet motor design, McGraw-Hill NewYork.

126

Page 143: The Design of a Micro-turbogenerator

Herrault, F. (2009), “Microfabricated air-turbine and heat-engine-driven permanent-magnet generators,” Ph.D. thesis, University of Toulouse.

Herrault, F., Arnold, D., Zana, I., Galle, P., and Allen, M. (2008), “High temperatureoperation of multi-watt, axial-flux, permanent-magnet microgenerators,” Sensorsand Actuators A: Physical, 148, 299–305.

Herrault, F., Yen, B., Ji, C., Spakovsky, S., and Lang, J. (2010), “Fabrication andPerformance of Silicon-Embedded Permanent-Magnet Microgenerators,” Journalof Microelectromechanical Systems, 19.

Holmes, A., Hong, G., and Pullen, K. (2005), “Axial-flux permanent magnet ma-chines for micropower generation,” Microelectromechanical Systems, Journal of,14, 54–62.

Honda Corporation (2001), “Honda EU1000i datasheet,” .

Jovanovic, S. (2008), “Design of a 50-Watt Air Supplied Turbogenerator,” Master’sthesis, Massachusetts Institute of Technology.

Lang, J. (2009), Multi-Wafer Rotating MEMS Machines: Turbines, Generators, andEngines, Springer.

Laughton, M. and Warne, D. (2003), Electrical engineer’s reference book, Newnes.

Lee, C., Arslan, S., and Frechette, L. (2008), “Design Principles and Measured Per-formance of Multistage Radial Flow Microturbomachinery at Low Reynolds Num-bers,” Journal of Fluids Engineering, 130, 111103.

Mattingly, J., of Aeronautics, A. I., and Astronautics (2006), Elements of propulsion:gas turbines and rockets, American Institute of Aeronautics and Astronautics.

Mehra, A., Zhang, X., Ayon, A., Waitz, I., Schmidt, M., and Spadaccini, C. (2002),“A six-wafer combustion system for a silicon micro gas turbine engine,” Microelec-tromechanical Systems, Journal of, 9, 517–527.

Narayan, S. and Valdez, T. (2008), “High-Energy Portable Fuel Cell Power Sources,”Electrochemical Society Interface, p. 41.

Nave, C. (2011), “HyperPhysics,” ”http://hyperphysics.phy-astr.gsu.edu/hbase/hph.html#hph”.

Osipov, I. L. (2008), “A Choice of Trailing Edge Diameter of Gas Turbine Blades,”Thermal Engineering, 55, 438–441.

Panasonic Corporation (2009), “Panasonic Lithium Ion CGR26650A datasheet,” .

127

Page 144: The Design of a Micro-turbogenerator

Peirs, J., Reynaerts, D., and Verplaetsen, F. (2003), “Development of an axial mi-croturbine for a portable gas turbine generator,” Journal of Micromechanics andMicroengineering, 13, S190.

Pelekies, S., Schuhmann, T., Gardner, W., Camacho, A., and Protz, J. (2010), “Fab-rication of a mechanically aligned single-wafer MEMS turbine with turbocharger,”in Proceedings of SPIE, vol. 7833, p. 783306.

Raisigel, H., Cugat, O., and Delamare, J. (2006), “Permanent magnet planar micro-generators,” Sensors and Actuators A: Physical, 130, 438–444.

Roberts, Q. D. and Denton, J. D. (1996), “Loss Production in the Wake of a Sim-ulated Subsonic Turbine Blade,” in International Gas Turbine and AeroengineCongress and Exhibition, Birmingham, UK, ASME Paper 96-GT-421.

Savoulides, N., Jacobson, S. A., Li, H., Ho, L., Khanna, R., Teo, C. J., Protz, J. M.,Wang, L., Ward, D., Schmidt, M. A., and Epstein, A. H. (2008), “Fabrication andTesting of a High-Speed Microscale Turbocharger,” Journal of Microelectrome-chanical Systems, 17.

Schlichting, H. (1968), Boundary-Layer Theory, McGraw Hill Companies, Inc., NewYork, 6 edn.

Shapiro, A. H. (1953), The Dynamics and Thermodynamics of Compressible FluidFlow, vol. 1, The Roland Press Company, New York.

Shepherd, D. G. (1956), The Principles of Turbomachinery, MacMillan PublishingCompany, New York.

Spadaccini, C., Zhang, X., Cadou, C., Miki, N., and Waitz, I. (2002), “Developmentof a catalytic silicon micro-combustor for hydrocarbon-fueled power MEMS,” inMicro Electro Mechanical Systems, 2002. The Fifteenth IEEE International Con-ference on, pp. 228–231, IEEE.

Spadaccini, C., Mehra, A., Lee, J., Zhang, X., Lukachko, S., and Waitz, I. (2003),“High Power Density Silicon Combustion Systems for Micro Gas Turbine Engines,”Journal of Engineering for Gas Turbines and Power, 125.

White, F. M. (1991), Viscous Fluid Flow, McGraw-Hill, Inc., New York, NY.

Yen, B., Lang, J., and Spakovszky, Z. (2008a), “A fully-integrated multi-wattpermanent-magnet turbine generator,” Ph.D. thesis, Massachusetts Institute ofTechnology.

Yen, B., Herrault, F., Hillman, K., Allen, M., Ehrich, F., Jacobson, S., Ji, C., Lang,J., Li, H., Spakovszky, Z., et al. (2008b), “Characterization of a fully-integratedpermanent-magnet turbine generator,” Proc. 8th PowerMEMS, pp. 121–124.

128