Linearly-Acting Variable-Reluctance Generator for Thermoacoustic Applications MASS, by CHUSETTS INSTJTULf OF T c' AUG 15 2014 Philip Clinton Knodel B.S., United States Air Force Academy (2012) LIBRARIES Submitted to the Department of Mechanical Engineering in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY June 2014 @ Massachusetts Institute of Technology 2014. All rights reserved. Signature redacted ...................... Department of Mechanical Engineering May 9, 2014 A uthor ... Signature redacted Certified by.... Signature redacted John G. Brisson Professor Thesis Supervisor .................... David E. Hardt Chairman, Department Committee on Graduate Theses Accepted by .... This work is sponsored by the Department of the Air Force under Air Force Contract #FA8721-05-C-0002. Opinions, interpretations, conclusions and recommendations are those of the author and are not necessarily endorsed by the United States Government. 2 Linearly-Acting Variable-Reluctance Generator for Thermoacoustic Applications by Philip Clinton Knodel Submitted to the Department of Mechanical Engineering on May 9, 2014, in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering Abstract Advances in battlefield equipment have created a demand for portable power systems with greater power density and more flexibility than current battery sources alone can provide. One potential solution lies in portable, high-frequency thermoacoustic engines, which can provide a battery recharge station or direct power supply for soldiers in-field, with the flexibility of operating on any quality heat source such as a butane heater. In this work, a linearly-acting variable-reluctance generator (VRG) is developed to act as a high-frequency (250 Hz) electroacoustic transducer and extract electric power from a proposed thermoacoustic engine design. A computational model of the thermoacoustic engine was developed using DeltaEC to determine the feasibility of the concept and the necessary characteristics of the transducer element. The unique requirements of the high-frequency thermoacoustic engine led to the design, optimization, fabrication, and testing of the VRG, which is designed to operate resonantly in the thermoacoustic system and supported by a self-pumping gas bearing system. The VRG is uniquely suited to efficiently convert small amplitude mechanical oscillations (5 mm in this work) into electric power. Linear and nonlinear saturation models of the transducer were developed to optimize the VRG design and predict transducer performance in the thermoacoustic engine. The accuracy of these models was established by comparing simulation results to static experimental data and dynamic experimental data taken at 50-60 Hz oscillation frequency, testing one half of the full transducer. Experimental testing resulted in a maximum power output of 5.74 Watts with an efficiency of 55.8%. The results led to the conclusion that the transducer would function as designed in the thermoacoustic engine. Recommendations for future work and guidelines for future development of the engine, transducer, and bearing system are provided based on the design and results presented in this work. Thesis Supervisor: John G. Brisson Title: Professor 3 "Man's flight through life is sustained by the power of his knowledge." -Austin 'Dusty' Miller 4 Acknowledgments I would like to extend a special thank you to Sumanth Kaushik for giving me this opportunity to work and study at MIT the past two years. I would also like to thank Lincoln Laboratory and the Advanced Concepts Committee for funding my studies and research. I have an amazing family and I have to thank them and especially my parents Bryan and Lael for their continued encouragement, especially through the toughest times. I would not be where I am today without their love and support even from afar. There are a number of other people who have been particularly important in my academic, personal and professional development. Professor John Brisson has been my research adviser the past two years and his mentorship and direction in my thesis work has been greatly appreciated. Professor Jeffrey Lang was also instrumental in my education and helping me come up with a research path while at MIT. I cannot thank them enough for their patience and instruction. The team that formed around the thermoacoustic project became invaluable to me. To my fellow graduate student, Claudio Hail, I thank you for your dedication to the project and everything you brought to the team. I wish you all the best in your academic studies and beyond. I also thank the UROP, Niharika Bhargava, for her assistance on the project. The Cryolab has been a great place to conduct research for the past two years and I would like to thank Doris Elsemiller, Michael Demaree, Paul Flinn, and Don Strahan for their insights and ability to make the lab run so smoothly. Also a thank you to my labmates Nick Roche, Martin Segado, Victoria Lee, and Melissa Ireland for making the lab an enjoyable place. Finally, I would like to thank my friends at MIT for making MIT a great place to work and Cambridge a fun place to live, and most importantly, I thank God who gives all things to all people including the strength to persevere and the will to look ever skyward. 5 DISCLAIMER: The views expressed in this article are those of the author and do not reflect the official policy or position of the United States Air Force, Department of Defense, or the U.S. Government. 6 Contents Abstract 3 Acknowledgements 5 M otivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 1.2 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 1.3 Standing and Traveling Wave Engines. . . . . . . . . . . . . . . . . 17 1.4 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 . . . . . . . . . . . . . . . . . . . . . . . . 25 . . . . . . . . . . . . . . . . . 31 . . . . . . . . . . . . . . . . 33 . . 1.4.1 Thermoacoustics 1.4.2 Electroacoustic Transduction . Thesis Overview . . . . . . . . . . . . Thermoacoustic Generator Description 34 2.1.1 DeltaEC Model . . . . . . . . . . . . . . . . . . . . . . . . 39 2.2 Variable Reluctance Generator Design . . . . . . . . . . . . . . . . 48 2.3 Gas Bearing Design . . . . . . . . . . . . . . . . . . . . . . . . . . 52 2.3.1 Design Iterations . . . . . . . . . . . . . . . . . . . . . . . 54 2.3.2 Final Gas Bearing Concept . . . . . . . . . . . . . . . . . 57 . . . . . . . . . . . . . . . . . . . . . . . . . . 63 2.4 Chapter Summary . . . . . . . . . . . . . . . . . . . . . 34 Engine Design . . . . . . . . . . . . . 2.1 Design of Variable Reluctance Generator 3.1 D esign . . . . . . . . . . . . . . . . . . . . 3 . 1.1 1.5 2 13 Introduction . 1 7 64 64 . . . . . . . . . . . . 69 3.2.1 Linear Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 3.2.2 Nonlinear Saturation Model . . . . . . . . . . . . . . . . . . . 75 3.2.3 Loss Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . 79 3.2.4 Flux Tube Analysis and Fringing . . . . . . . . . . . . . . . . 82 3.3 Power Electronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85 3.4 Design Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 3.5 VRG Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 . . . . . . . . . . . . . . . . . . . . . 92 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 3.2 3.6 VRG Model .. .. . . .. .. ...... 3.5.1 Laminated Components 3.5.2 C oil Chapter Summary .... 99 4 Experimental Design Verification Saturation Characterization 4.2 Experiment Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 4.3 Dynamic Experimental Results . . . . . . . . . . . . . . . . . . . . . 112 Frequency Scaling and Losses . . . . . . . . . . . . . . . . . . 125 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 4.3.1 4.4 5 . . . . . . . . . . . . ........... 99 4.1 Chapter Summary 130 Conclusions and Future Work 5.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 5.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 . .... . . . . . . . . . 133 5.2.1 M odeling 5.2.2 System and Fabrication Improvements A Flux Tube Analysis 134 B Model Code and Experimental Data 137 C Matlab Code for VRG 141 D VRG Component Drawings 146 8 List of Figures 1-1 Sondhauss-Tube . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 1-2 R ijke-Tube . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 1-3 Standing Wave Engine Description . . . . . . . . . . . . . . . . . . . 18 1-4 Brayton Cycle T-s Diagram and Modified T-s Diagram . . . . . . . . 20 1-5 Effect of Compresion Ratio and TH on the Brayton Cycle . . . . . . . 22 1-6 Traveling Wave Engine . . . . . . . . . . . . . . . . . . . . . . . . . .. 24 1-7 Swift/Backhaus Traveling Wave Engine . . . . . . . . . . . . . . . . . 26 1-8 NASA HEPS Engine . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 1-9 Yazaki Looped Tube Thermoacoustic Engine . . . . . . . . . . . . . . 28 1-10 Score Looped Tube Engine . . . . . . . . . . . . . . . . . . . . . . . . 29 . . . . . . . . . . . . . . . 30 2-1 Proposed Thermoacoustic Engine Schematic . . . . . . . . . . . . . . 35 2-2 Double Helmholtz Resonantor . . . . . . . . . . . . . . . . . . . . . . 36 2-3 Resonant Engine with Nonlinear Transduction Proposed by Swift . . 36 2-4 DeltaEC Schematic . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 2-5 Pressure, Volume Flow Rate, Acoustic Power and Total Power Profiles 1-11 Aster Thermoacoustic Multi-stage Engine in the Thermoacoustic Engine . . . . . . . . . . . . . . . . . . . . . . 42 2-6 Stack Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 2-7 Effect of Stack Location . . . . . . . . . . . . . . . . . . . . . . . . . 45 2-8 Geometry Optimization . . . . . . . . . . . . . . . . . . . . . . . . . 46 2-9 Effect of Stack Gap Size . . . . . . . . . . . . . . . . . . . . . . . . . 47 2-10 Radial vs Axial Air Gap Linear Alternator . . . . . . . . . . . . . . . 51 9 52 ............................ 2-11 VRG Concept ...... 53 2-13 Gas Diode Based Bearing Design . . . . . . . . . . . . . . . . 55 2-14 Piezoelectric Gas Diode Pressures . . . . . . . . . . . . . . . . 56 2-15 Unsuccessful Sliding Piston Check Valve Gas Bearing . . . . . 58 2-16 Schematic and Pressure Profiles of Gas Bearing Breakdown . . 59 . . . . . . . . . . 60 2-18 Relative Piston Position for Charging and Discharging Plenums 60 2-19 Gas Bearing Restoring Force . . . . . . . . . . . . 62 . . . . . . . . . . . 65 3-2 VRG Design Diagram . . . . . . . . . . . . . . . . . . . . . . . . . 65 3-3 VRG Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 3-4 VRG Assembly Including Piston and Piston Cap . . . . . . . . . . . 68 3-5 Magnetic Circuit Diagram ............... . . . . . . . . . . . 69 3-6 Inductance versus Air Gap Length . . . . . . . . . . . . . . . . . . 72 3-7 Cyclic Inductance Motoring Versus Generating . . . . . . . . . . . . 73 3-8 Ideal Flux Linkage-Current Profiles . . . . . . . . . . . . . . . . . . 74 3-9 Flux Linkage-Current Plots with Saturation . . . . . . . . . . . . . 76 . . . . . . . . . . . . . . . . . . . . . . . . . . 77 3-11 Potential Current Waveforms and Cyclic Flux Linkage Current Loops 78 3-12 Minor Hysteresis Loop . . . . . . . . . . . . . . . . . . . . . . . . . 81 3-13 Magnetic Circuit Diagram Including Fringing and Leakage Fields . 83 3-14 Flux Tube Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 . . . . . . . . . . . . . . . . . . . . . . . . 86 3-17 Drive Circuitry PCB Schematic . . . . . . . . . . . . . . . . . . . . 87 3-18 Potential Excitation Stages . . . . . . . . . . . . . . . . . . . . . . . 88 3-19 Winding Height Dimension Optimization . . . . . . . . . . . . . . . 90 3-20 Optimimum Air Gap Geometry . . . . . . . . . . . . . . . . . . . . 91 3-21 Laminated Stator and Piston Components . . . . . . . . . . . . . . 92 . . . . . . . . . . . 3-16 Drive Circuit Schematic . . . 3-10 Magnetic Force Plot . Stator Design . . . . . . . . . . . . . . . . . . . . 3-1 . . 2-17 Schematic of Proposed Gas Bearing System . . . . . . 2-12 Gas Bearing Voltage Divider Analogy . . . . . . . . . . . . . . 10 3-22 Materials and Jig Used for Stator Assembly . . . . . . . . . . . . . . 93 3-23 Epoxy Cure Bake Assembly . . . . . . . . . . . . . . . . . . . . . . . 94 3-24 Stator Assembly with Brackets 95 . . . . . . . . . . . . . . . . . . . . . 3-25 Assembly of Piston Laminated Components . . . . . . . . . . . . . . 95 3-26 Coil Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97 3-27 Stator Assembly with One Phase Coil . . . . . . . . . . . . . . . . . . 97 4-1 Static Experiment Air Gap Shims . . . . . . . . . . . . . . . . . . . . 100 4-2 Circuit Diagram for Generating Flux Linkage-Current Plots 101 4-3 Current and Voltage Profiles for Generating Flux Linkage-Current Plots102 4-4 Flux Linkage-Current Plots from Experiment Data . . . . . . . . . . 103 4-5 Correction to Fringing Fluxes Based on Experimental Data . . . . . . 104 4-6 Comparison of Approximate B-H Curves to Published Data . . . . . 106 4-7 Experimental Setup Schematic . . . . . . . . . . . . . . . . . . . . . . 110 4-8 Dynamic Test of VRG . . . . . . . . . . . . . . . . . . . . . . . . . . 110 4-9 Experiment Drive Circuitry and Power Dissipation . . . . . . . . . . 112 4-10 Predicted Efficiency and Power Curves . . . . . . . . . . . . . . . . . 113 4-11 Predicted versus Measured Flux Linkage-Current Loops . . . . . . . . 116 4-12 50 Hertz Data Point for Power Output . . . . . . . . . . . . . . . . . 117 4-13 50 Hertz Data Point for Efficiency . . . . . . . . . . . . . . . . . . . . 118 4-14 60 Hertz Data Point for Power Output . . . . . . . . . . . . . . . . . 120 4-15 60 Hertz Data Point for Efficiency . . . . . . . . . . . . . . . . . . . . 121 4-16 60 Hertz Comparison of Model to Data Power Output . . . . . . . . . 122 4-17 60 Hertz Comparison of Model to Data Efficiency . . . . . . . . . . . 123 4-18 Effect of Minimum Air Gap on Efficiency and Power 4-19 Operating Frequency Effect on VRG Losses . . . . . . . . . . . . . . 124 . . . . . . . . . . . . . . 126 4-20 Frequency Scaled True Power Output and Efficiency . . . . . . . . . . 128 A-1 Flux Tube Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135 11 2.1 . List of Tables Thermoacousic engine parameters . . . . . . . . . . . . . . . . . . . 40 2.2 Gas Bearing Operating Parameters . . . . . . . . . . . . . . . . . . . 61 3.1 VRG final design dimensions . . . . . . . . . . . . . . . . . . . . . . . 67 3.2 Drive Circuitry PCB Components . . . . . . . . . . . . . . . . . . . . . 86 4.1 Estimated VRG Core Loss . . . . . . . . . . . . . . . . . . . . . . . . 127 12 Chapter 1 Introduction 1.1 Motivation Soldier pack weight has become an increasingly important issue as technological advances in battlefield equipment have broadened the capabilities of the war-fighter, but also increased the electrical power demands in remote locations. With these advances, soldiers have been asked to carry more and more weight in the form of batteries, in addition to all of their other gear. Currently, soldiers in the field carry between 10 and 16 pounds of batteries, which is roughly 20 percent of their pack weight. The rate at which batteries are used is also an alarming figure. An infantry company, roughly 150 soldiers, will use 6,600 batteries amounting to 1,400 pounds in 72 hours of operation. The army reports that yearly spending on batteries for a battalion is roughly $150,000 and is second in magnitude only to munitions. There are a number of ongoing research efforts aimed at solving the portability of electrical power for soldiers. A few of these attempts include portable photovoltaics known as the Rucksack Enhanced Portable Power System (REPPS) and fuel cell research being conducted by CERDEC and DARPA. However, both of these systems have serious limitations. The weight of the REPPS system is in the range of 4-7 pounds and produces 55 Watts in optimal conditions. The power density of this system is not unreasonable, but the photovoltaics systems themselves are limited to day use, in an open/exposed area, and are climate dependent. 13 These systems should not be discounted, but must be augmented if the soldier is expected to rely on rechargeable batteries. Fuel cell research is a promising technology for replacing conventional batteries. The energy density of hydrogen used in fuel cells is nearly four orders of magnitude greater than the best batteries to date. However, current fuel cells require a highly purified hydrogen source to operate. This is the current technological shortcoming, particularly for operation as a deployed power source. Further research will be required before fuel cells become a viable alternative to conventional batteries for portable soldier power. An alternative to these technologies is thermoacoustic engine based technology. Thermoacoustic engines have many desirable attributes because the engine operates with the application of any quality heat source and requires few to no moving components. These attributes make the engine both flexible and robust as a portable power source. The challenges associated with applying thermoacoustic technology to meet portable power demands are scaling down existing technology, increasing the power density in terms of watts per pound, and efficiently coupling acoustic to electric power. This collaborative program between Lincoln Laboratory and Massachusetts Institute of Technology (MIT) aims to make advances toward solving each of these challenges. This work details the design of the Portable ThermoAcoustic Generator (PTAG), and to greater detail, the acoustic-to-electric transducer component within the thermoacoustic engine. 1.2 Background Thermoacoustic technology has been applied to a number of different applications including: refrigeration, cryogenics, natural gas liquefaction, space power, remote power generation, waste heat recovery, cooling of microelectronics, and solar cooling of homes. Some of the most significant and foundational research in thermoacoustics was led by Greg Swift at Los Alamos National Laboratory. In the scope of thermoacoustics, there are two forms of engines: standing wave 14 Figure 1-1: The Sondhauss-tube is the earliest investigation of a thermoacoustic type device. Figure originally from Rott [19]. and traveling wave. At the most basic level, both standing and traveling wave engines operate by correctly phasing heat transfer to/from a fluid that is also undergoing the compression and rarefaction of acoustic type waves. A resonant system and a temperature gradient are used to correctly phase the heat transfer in all thermoacoustic devices, where the temperature gradient is maintained along the same direction as bulk fluid motion at a specific location(s) in the resonant system. Different methods of properly phasing the heat transfer and setting the resonance of the system are what separates standing wave engines and traveling wave engines, and has led to the design of numerous forms of thermoacoustics devices, several of which are discussed in Section 1.4. The details of the difference between the standing wave and traveling wave engines is discussed in more detail is Section 1.3. The thermoacoustic phenomenon was first observed by glass blowers when a hot bulb was placed against a cold tube as shown in Figure 1-1. In certain dimensions and temperature differences between the hot bulb and cold tube, this interaction would cause the spontaneous generation of pressure oscillations (sound) from the end of the cold tube, later deemed a Sondhauss-tube after the first experimenter [19]. In this configuration, the temperature gradient was established by conduction from the hot bulb to the cold tube, and the specific geometries of the bulb and tube are what set the acoustic resonance of the system. The sound would quickly fade from these tubes as the temperature gradient diminished and equilibrium between the bulb and tube 15 Sound v~m ,pip(geeraon Gauze heated Jkjnveoflow Figure 1-2: The Rijke-tube, which was the first thermoacoustic engine to use a stack (the steel gauze) to maintain the temperature gradient. Figure originally from Matveev [12]. was established. The Sondhauss-tube was followed by the Rijke tube in which steel gauze was used to maintain the temperature gradient as shown in Figure 1-2. The use of a porous matrix to establish the temperature gradient became known as a stack or regenerator in thermoacoustic devices. "Stack" is typically reserved for standing wave engines, while "regenerator" is used to describe the porous medium of traveling wave engines, which is discussed further in the following section. The acoustic resonance of the Rijke tube, like that of the Sondhauss-tube, was maintained by the geometry of the acoustic wave enclosure, labeled "vertical pipe" in the Rijke tube figure. Lord Rayleigh provided the qualitative description of the engine in his 1877 text, The Theory of Sound. Lord Rayleigh writes for the description of this engine [16] For the sake of simplicity, a simple tube, hot at the closed end and getting gradually cooler towards the open end, may be considered... If heat be given to the air at the moment of greatest condensation, or be taken from it at the moment of greatest rarefaction, the vibration is encouraged. Following Lord Rayleigh's qualitative description, Nikolaus Rott developed the foundations of thermoacoustics, which are used for the analysis of thermoacoustic 16 engines today. Rott's work is based on linear acoustic theory, which Swift describes in great detail in his Thermoacoustics text [24]. 1.3 Standing and Traveling Wave Engines Standing Wave Engine An example of a simple standing wave engine is shown in Figure 1-3 part (a) with a resonant tube, closed at both ends, and a stack centered at a distance L, from the left tube wall. Acoustically, this configuration is known as a half-wavelength resonator. The acoustic wavelength being given by the equation a A= - (1.1) f where A is the wavelength, a is the speed of sound in the working fluid, and f is the frequency of oscillation. The resonant frequency of the half wavelength resonantor is set by the speed of sound in the fluid and the length of the resonant chamber, L,. The half wavelength is formed by a rightward traveling wave and a leftward traveling wave, which upon reflection off of the ends of the tube forms a half of a single standing wave with a velocity node and a pressure antinode at each tube end, as shown in part (b) of Figure 1-3. Inherent in a standing wave is a 90 degree phase shift between the pressure and velocity oscillations, which is particularly important when considering the thermodynamics of the gas parcels within the stack. With the pressure and velocity oscillations 90 degrees out of phase, the bulk gas motion is in phase with the pressure oscillations such that gas parcels within the stack are at their leftmost extreme (hot side) when the pressure is high in the stack and at their rightmost extreme (cold side) when the pressure is low in the stack. To meet Rayleigh's criterion for the standing wave engine, heat transfer to the gas must occur at high pressure and heat transfer from the gas must occur at low pressure as shown in part (c) of Figure 1-3. To achieve this phasing, poor thermal coupling between the stack and gas is required, such that the gas experiences adiabatic compression and 17 Lc LA/2 = A/20 =L 0Resonant Tube Gas Particle Motion .... Figure Part (c) Stack Cold Heat Exchanger Hot Heat Exchanger (a) PMean + i ~~ Volume Flow --- - - --~- --~ Rate +lull - 7e'ur7e I -'phl -lU1l Volume Flow Rat--'--.. - -' !ressure_- --- (b) O, =o*, Oe = O,= 90*, eu= 90* * Brayton Cycle P-V Diagram C B 0 - I -- -- - AS=O Expansion -Ipl - - - -- C AS=O compression -- ~~~~~~~~~ --- ~~ A D Qut < _ --------- _ __ A I . --- cr Volume (V) (C) (d) Figure 1-3: Shown in (a) is a simple example of a standing wave engine in a half wavelength resonant tube with the thermoacoustic stack located a length L, from the left wall. (b) The phasing of a half wavelength standing wave is shown. The phasing of the wave is necessary for understanding (c) the relative phasing of heat transfer, pressure and particle motion in the stack. (d) The relative phasing of the thermodynamic interactions of the gas parcels and porous stack produces a thermodynamic cycle within each gas parcel nearly identical to the Brayton cycle shown. 18 expansion and heat transfer is limited to isobaric conditions at the extremes of the gas parcel motion. Imperfect thermal contact between the gas and the solid walls of the stack is accomplished by making the stack gap spacing or pore hydraulic radius, rh, perpendicular to the bulk fluid motion approximately the size of the thermal penetration depth [2]. The thermal penetration depth is given by 6k 2k WPCp (1.2) where k is the thermal conductivity of the gas, p is the gas density, c, is the specific heat of the gas, and w is the angular frequency of the bulk gas oscillation. With the stack plates, honeycomb or other porous material spaced in this fashion, the majority of heat transfer occurs at the extremes of fluid motion and Raleigh's criterion is met. The adiabatic compression, isobaric heat transfer to the fluid, adiabatic expansion, and isobaric heat heat transfer from the fluid is characteristic of the Brayton cycle, which is shown on a pressure-volume diagram in part (d) of Figure 1-3. The actual motion of the gas parcels is sinusoidal, and thus the true P-V diagram is not discrete thermodynamic steps, but the idealized form is useful for both conceptual understanding, and examining the limits of the standing wave engine. The standing wave engine cannot exceed the Carnot efficiency, 77c = 1 - -- TH (1.3) In addition to this limit, standing wave engines are inherently inefficient due to the imperfect thermal contact between the stack and gas. As the contact between the gas and stack -increases, such that the lateral spacing in the stack decreases below the thermal penetration depth of the gas, the adiabatic compression and expansion assumptions are no longer valid and Rayleigh's criterion is no longer met. The effect of increasing the thermal contact becomes particularly clear when the cycle is plotted on a T-s diagram as in Figure 1-4, where the increase in entropy during compression and decrease in entropy during expansion is caused by heat transfer to and from the 19 Ideal Brayton Cycle T-s Diagram Modified Brayton Cycle T-s Diagram C C AS=0 4- Expansion 0. E a) 40. E D B D B a) AS=O Compression A A Entropy (s) Entropy (s) (b) (a) Figure 1-4: (a) The ideal Brayton T-s digram is shown. (b) In contrast to the ideal Brayton cycle, increasing the thermal contact between the gas and the stack causes non-adiabatic compression and expansion. In the extreme of perfect thermal contact, no asymmetry exists and no net work is produced by the cycle. gas respectively. As the contact between the gas and stack becomes perfect, there no longer exists thermodynamic asymmetry, and the temperature and entropy of the gas follow the dotted black line in the figure with no net work transferred to the gas over the course of an oscillation. Therefore, the imperfect thermal contact is fundamental to the standing wave engine. Although imperfect thermal contact is required for a standing wave engine to operate, the efficiency can still theoretically approach the Carnot efficiency. In the limit that the temperature gradient of the stack exactly matches the temperature increase of the gas due to adiabatic compression, the heat transfer to the gas parcel occurs over an infinitesimally small temperature difference, which indicates no entropy generation. However, with an infinitesimally small temperature difference, the heat transfer is also infinitesimally small. This limit is known as the critical temperature gradient and is expressed as VTcrt Pm - I P.cA|1| 20 (1.4) where 1p, I and IjI are the magnitude of the pressure and displacement oscillations respectively, and Pm is the mean density of the gas. Therefore, the standing wave engine can only approach Carnot efficiency in the limit that no acoustic power is generated. This is of little practical use, so the heat transfer must occur over a large temperature difference to produce meaningful acoustic power. Entropy generation increases with the square of the temperature difference, which indicates that the standing wave engine must operate with an efficiency significantly below the Carnot efficiency. Due to this inherent entropy generation, the actual efficiency of standing wave engines to date is approximately 10-25% of Carnot's efficiency [24]. Other prime movers operating from the Brayton cycle largely overcome the poor efficiency characteristics of the cycle by increasing the compression ratio P2 rP PL (1.5) where PL is the minimum pressure before compression and P2 is the pressure of the working fluid immediately following compression. By increasing the compression ratio, the temperature difference, AT, during heat transfer to the fluid is minimized as shown in Figure 1-5. In part (a) of the figure, three Brayton cycles are shown with increasing compression ratios. The heat transfer to the fluid, which occurs from points B to C in the figure, occurs with a smaller temperature difference between the gas and the the hot side temperature at larger compression ratios. For thermoacoustic engines, the compression ratio is given in terms of a drive ratio defined as Dr = PM (1.6) where Pm is the average pressure of the working fluid. The standing wave engine drive ratio is limited to approximately 0.1 before the acoustics become significantly nonlinear and the efficiency drops. The drop in efficiency has been suggested to occur due to increased power dissipation in acoustic harmonics formed within the resonant system[23]. Therefore, the compression ratio for the standing wave engine is limited to approximately 1.3, which is significantly smaller than typical Brayton cycle engines 21 Brayton Cycle T-s Diagram Brayton Cycle T-s Diagram (Varying Compression Ratio) (Varying TH) 4 TH rp =8 C3 C TH - r,=15 .-------- ---- ----- ---- ----- ------ 3 AT D3 C2 TH CL M 2 B E E a) 9 C1 TH1 T- C1T D2 D1 A Entropy (s) Entropy (s) (b) (a) Figure 1-5: (a)The three Brayton cycles shown are for increasing compression ratios, rp. The larger compression ratios experience heat transfer (points B-C for the middle cycle shown) across a smaller temperature difference, AT, thus limiting entropy generation. (b) The only way to decrease entropy generation in the standing wave engine is to decrease the temperature at which heat transfer occurs, which comes at the expense of power output (the area enclosed in the cycle loop). operating with compression ratios of up to 30-40. The smaller compression ratio of the standing wave engine necessarily leads to heat transfer across a larger temperature difference for all meaningful acoustic power outputs. This is the primary limitation of the standing wave engine and has been the impetus for research in the area of traveling wave thermoacoustic engines. Traveling Wave Engine Traveling wave thermoacoustic engines rely on a feedback path, such as a looped tube, to provide acoustic power to the cold side of the regenerator, which is subsequently amplified while traveling through the regenerator toward the hot side. A simple example of this type of engine is shown in Figure 1-6, where the standing wave stack has been replaced by the traveling wave regenerator. In contrast to a stack, the regenerator is made such that the gas remains in intimate thermal contact with the 22 surrounding porous solid (rh << 6 k). While this configuration produces no power output in a standing wave engine, the excellent thermal contact is necessary for the operation of the traveling wave engine because of the difference in pressure and velocity phasing between a standing wave and a traveling wave. Unlike the standing wave engine, a traveling wave engine has pressure and velocity oscillations that are nearly in phase in the regenerator. The pressure is therefore high while the gas moves toward the hot heat exchanger and low while it moves toward the cold heat exchanger. Examining the behavior of a single gas parcel within the regenerator, as shown in part (b) of Figure 1-6, the gas is: 1. Compressed nearly isothermally while at its leftmost extreme position. 2. Moves laterally toward the hot side while accepting heat transfer from the porous regenerator, which was transferred there a half cycle earlier. 3. Expands isothermally given the excellent thermal contact at the parcels rightmost extreme. 4. Moves laterally toward the cold end while conducting excess thermal energy to the porous regenerator material. First noted by Ceperly, these cyclic conditions for the individual gas parcels is nearly identical to the Stirling cycle, which is shown as a p-V diagram is part (c) of Figure 16. The actual motion of the particles as in the standing wave engine is sinusoidal and thus the discrete cycle shown is an approximation of the true cycle of the individual gas parcels, but is useful for both developing understanding and analyzing the engine thermodynamically. The limit of the traveling wave engine is also set by Carnot's efficiency, but the excellent thermal contact leads to (ideally) reversible heat transfer. These conditions are not limited to a certain amplitude of pressure oscillation, and therefore meaningful acoustic power can be generated more efficiently than in a standing wave engine, with the highest efficiency to date reaching 41% of Carnot from thermal to acoustic power [3]. 23 Figure Part (b) Acoustic Power In Acoustic Power Out __ Cold Heat Regenerator Hot Heat Exchanger Exchanger Resonant Looped Tube (a) Isothermal Compression Regen. Heat In (+ P , +| UI) 'B II K 'loo*p,* rh«&5k OUt A 6k C Regen. Heat Out (-IPi, -IU, 1 Isothermal Expansion ) (b) Stirling Cycle T-s Diagram Stirling Cycle p-V Diagram T P D T+ .C--------C 0. B _0 12i 0 ---.. -1pil ------------- CI --A A c B C----- ~~~ ~-- S V (d) (c) Figure 1-6: (a) An example of a traveling wave engine in a looped tube is shown. (b) The gas parcels within the regenerator undergo the cycle shown, which effectively amplifies acoustic power as the wave travels from the cold side to the hot side. (c) The phasing shown in the second part of this figure is essentially a Stirling cycle shown on a p-V diagram and (d) on a T-s diagram. The heat transfer occurs across (ideally) no temperature gradient. 24 Although the efficiency of the traveling wave engine can be greater than the standing wave engine, the traveling wave engine is far more complex from a design standpoint. Additional limitations include viscous losses in the small channels of the regenerator, time averaged DC streaming around the loop (Gideon streaming), and extracting power from a traveling acoustic wave. The regenerator viscous losses and Gideon streaming are mentioned here only for completeness, but extracting power from the acoustic waves is discussed further in Section 1.4.2 and Section 2.2. 1.4 1.4.1 Literature Review Thermoacoustics The Condensed Matter and Thermal Physics Group at Los Alamos National Laboratory conducted research in a number of areas for thermoacoustic engines, including both standing and traveling wave type engines as well as combinations of the two [11] [23]. The most well known of Swift's engines is his "traveling wave engine," which is actually a standing wave engine with a looped gas RLC circuit in order to shift the pressure and velocity nodes to achieve traveling wave phasing inside the regenerator [3]. A schematic diagram of the engine is shown in Figure 1-7. One of the characteristics of engines of this type is that the acoustic and thermal components are significantly shorter than an acoustic wavelength. This is what allows the components to be treated as lumped circuit elements and analyzed using the acoustically analogous electric circuit parameters. Examining Swift's traveling wave engine, the resonator establishes a half wavelength resonance in the engine, but the loop at the left end, referred to as a torus, shifts the pressure-velocity phasing such that they are in phase within the regenerator. The traveling wave condition is met only at one specific location in the acoustic circuit, while the rest of the engine is, acoustically, a standing wave. This engine was ultimately capable of converting heat into acoustic power with an efficiency of 41% of Carnot and no moving components [3]. However, the engine 25 Mlain ambient heat exchanger Regenerator [lot heat exchanger Secondary ambient heat exchanger and flow Feedback inertancenr 20cm ie. naor junm To rsonAtor o M %1 P Variable acoustic load P We,, Resonator P Figure 1-7: This figure shows the Swift/Backhaus traveling wave engine, which reached a thermal efficiency of 30%. The engine operated at 80 Hz and produced 71OW of acoustic power. Figure adapted from [3]. did not produce any electrical power. Without an electroacoustic conversion mechanism this engine configuration is only useful for directly driving other systems that require acoustic power, such as a thermoacoustic refrigeration cycle. Combinations of thermoacoustic engines and refrigerators have been used in a number of application areas, including natural gas liquefaction [26], pulse-tube refrigeration [24], and microelectronics cooling [27]. Building off of Swift's traveling wave design, a similar engine configuration was investigated by NASA as a spacecraft radioisotope power source [15]. This engine, known as the High Efficiency Power Source (HEPS), was scaled down significantly in size and used a mechanical, instead of an acoustic, resonator for both compactness and to provide an acoustic-to-electric transducer. As shown in Figure 1-8, the same RLC phasing was used around the torus, but instead of the acoustic resonator the loop was attached to twin opposed linear alternators. These alternators were connected at the point labeled "Alternator Interface" on the diagram, and oscillated in a direction 26 Flat Plate Hot Heat Exchanger Thermal Buffer Tube egenerator R ject HX it Pump 00 omplinCe o Altrnator Inteface 0 0 0 0 0 0 0 0 0 II nertance -ine Figure 1-8: This figure shows the High Efficiency Power Source (HEPS) system, which was designed for operation on a spacecraft with a radioisotope heat source. The engine reached a thermal efficiency of 18%, operated at 120 Hz, and produced 58 Watts of electric power. The power density for the engine was 8.3 W/kg. Figure adapted from [15]. perpendicular to the loop as it is depicted in the figure. However, reports on the HEPS engine detail significant issues acoustically matching these linear alternators. With the torus and mechanical resonance, the HEPS engine was capable of producing 58 Watts of electric power with a thermal efficiency of 18%. The research undertaken in the current project is largely an extension of the HEPS engine, with more emphasis on robustness, higher power density, and alternative geometries, and less emphasis on high efficiency. More recently, a large collaborative effort in Europe has been undertaken by SCORE (Stove for Cooking, Refrigeration and Electricity) in the UK, the FACT Foundation, and Aster Thermoacoustics in the Netherlands to develop an electricity producing thermoacoustic generator (TAG) for use with cooking stoves in develop- 27 Stack LDV Glass tube e e aa by t P ressu re tra nsd T TC PhotomultiplierK Heat exchangers ,=0 or 1 uce rs A .......- Looped tube Figure 1-9: This figure shows te experimental setup of Yazaki et al to produce traveling waves in a looped tube. Figure taken from [32]. ing countries. Their initial efforts were to develop a standing wave engine, but were turned away by the pressure amplitudes and displacements necessary for the linear alternator [10]. This led to the development of what can be deemed a pure trav- eling wave engine. The pure traveling wave engine was first investigated by Yazaki et al, building off the work done by Ceperley [32], where the regenerator is placed in a looped tube as shown in Figure 1-9, and similar to the traveling wave example discussed in Section 1.3. While Yazaki's looped tube engine was able to spontaneously produce acoustic oscillations, the viscous losses in the regenerator were significant, due to large acoustic velocities, owing to very low acoustic impedance in the regenerator. There is a bal- ance in terms of the regenerator impedance for traveling wave engines, because lower impedance means higher acoustic velocities. Larger acoustic velocities can increase the amplification of acoustic power in the regenerator but this subsequently increases viscous losses as well [33]. The SCORE engines are an advance over that of Yazaki's looped tube engine, where the impedance in the regenerator is carefully controlled. Both the placement of the regenerator in the loop and an increase in the cross-sectional area of the loop are used to increase the pressure oscillations within the regenerator while decreasing 28 P2 it COk He T4 T3 Ragenwratr T2 T lEXcr (CyX) Hot Hoe Exco Aftmtor w ih c(HH)o ~ftFeedback p*e Figure 1-10: This figure shows the design and experimental setup of a looped traveling wave thermoacoustic engine with a loudspeaker acoustic-to-electric transducer developed by the SCORE initiative. The engine produced 10.5 watts of acoustic power with an efficiency of 1.93%. The acoustic-to-electric conversion efficiency was 52.5%. Figure taken from [1]. the velocity. The reduction in velocity helps to mitigate the viscous losses which dominated Yazaki's engine. The thermoacoustic traveling wave in the SCORE engine is coupled with a loudspeaker operating in reverse to convert acoustic to electric power as shown in Figure 1-10. The low-impedance loudspeakers are used to match the traveling acoustic wave [1]. However, to date, these eni a very limited in their efficiency. This low efficiency is due to both the near atmospheric mean operating pressures, and the low transduction efficiency of loudspeakers, converting at roughly 50% acoustic to electric compared to typical 80-90% of higher mass, higher cost linear alternators. Other methods for power conversion are under development, which will be discussed in greater detail in section 1.4.2. Both the mean pressure limitation and the use of loudspeakers for transduction are limits associated with the low-cost objectives of the SCORE project and are not inherent limitations in the thermoacoustic engine. For the application of a portable power source, these looped tube engines are still not feasible due to the large acoustic wavelengths and subsequent lengths of tubing required for the feedback path. Reduction in the size of the system is possible at higher 29 load #1load #4 24 load # # SHeat in at Thigh C]Heat out at T,, load 7 Acoustic loop power Figure 1-11: This figure shows the design of the waste heat recovery system developed by de Blok for a paper manufacturing plant. The engine produced 1.64 kW of acoustic power with an efficiency of 38% of Carnot. The engine was not used with the linear alternators because they were not correctly designed with the integral system. Figure taken from [1]. frequencies, but at higher frequencies the mass of the electroacoustic transducer acts to lower the frequency. This paradox is discussed further in Section 2.1. Kees de Blok, the founder of Aster Thermoacoustics, has generated a number of unique system geometries and has developed the use of multiple stages to lower the oscillation onset temperature [4]. The primary effort of de Blok has been to develop alternate feedback geometries that do not require resonators because of the large acoustic losses they generate. This has led to a similar looped tube concept as what is currently being investigated by SCORE. One of the largest projects completed by Aster Thermoacoustics was for waste heat recovery using a multistage pure traveling wave engine shown in Figure 1-11. The engine achieved an onset temperature differ- ence of only 45K, but the engine also experienced significant issues with the 8 twin opposed 1.25 kW linear alternators. These 4 sets of linear alternators are shown as the acoustic load elements in the figure. Due to the failure of the linear alternators, the engine was tested only for acoustic power production and not electrical power production. The works described heretofore do not constitute the whole body of literature, es- 30 pecially when considering the large number of simulation based papers and proposals, as well as literature that this researcher may be unaware of in coming to understand thermoacoustics and the state of the art (SOA). Design choices for the current project were based on optimizing the system for power density, robustness, and compactness, while attempting to incorporate the lessons learned and design guidelines of previous researchers. 1.4.2 Electroacoustic Transduction A thermoacoustic engine that produces acoustic power does not solve the objective of creating a portable power source for soldiers. A common failure/limitation of the thermoacoustic engines described in the previous section is the electroacoustic transducer. The primary issue found in the literature is the poor coupling between the acoustic waves and the transducer in terms of both the local pressures/displacements at the interface and in coupling of resonance between the acoustics and transducer suspension. For this reason, this project focused on building a thermoacoustic engine concept around the transduction mechanism. The transducer is the focus of this thesis, but the full engine concept is described in Chapter 2. The primary incentive for research into thermoacoutic engines is their potential for long lifetime and low maintenance due to removal of moving components. Swift made an attempt to completely remove oscillating components through magnetohydrodynamic [14] transduction. Unfortunately, the magnitude of the temperatures required to make this feasible, as well as the requirement to use an electrically conducting fluid, ruled this out for further investigation in this research. Additionally, the efficiency for the engine described by Swift and Migliori was less than 2%. Another method for transduction that has been investigated is direct coupling to piezoelectric material [13] [20] [27]. However, there is an impedance mismatch between the relatively compliant acoustic waves and the stiff piezoelectric material. Placing this material on a diaphragm can increase the compliance marginally, but typical transduction efficiencies to date have been less than 10% [13] [20]. This is not a realistic option when combined with a thermoacoustic engine because the thermal31 to-electric efficiency would be in the low percents. This efficiency was considered too low to be a practical portable power source and was not pursued further in this research. One of the other significant developments by de Blok is the concept of using a bidirectional turbine for converting traveling wave acoustic power to electrical power [5]. This concept appears to hold a significant amount of merit, and may, in the case of this research, be applied to scaling down the size of traveling wave technology. However, the width of the blades must be smaller than the displacement amplitude of the gas parcels, and therefore, there is a limit to the frequency at which these turbines could be applied. Ultimately, the engine described herein relies on the resonant mass of the piston and gas springs; and therefore, a bidirectional turbine, while an interesting idea, was not applied to this research. The SCORE project places a primary emphasis on the cost of the thermoacoustic waste heat engine. Therefore, their investigation has been limited to loudspeaker type transducers for converting acoustic to electric power [1] [5] [33]. Unfortunately, these transducers are 30-40% less efficient than the more costly linear alternators, can withstand only small pressure differentials across the thin diaphragm, have a low power density, and have reliability problems associated with the flexible connectors [17]. For example, in one of the SCORE looped tube engines, the conversion efficiency from acoustic to electric was 52.5% to produce 10.5 watts of electricity using a 0.0132 mr2 (20.5 in2 ) diaphragm [1]. The proposed cross sectional area for the engine in this research is 0.00456 m 2 (7 in2 ). These design issues are exacerbated in standing wave type engines as discussed in [10], and therefore, the loudspeaker type transducer was not used in this research. When searching for alternative transduction methods, the concept of oscillating capacitor plates was also investigated. However, an analysis showed that thousands of capacitive plates would be required to generate meaningful forces with an assumed breakdown voltage of 106 V/m. This was considered an unrealistic option for the purposes of power transduction. Other linear alternator designs are moving magnet and moving iron type linear 32 alternators. Typically, these type of transducers are more efficient, higher mass, and have a radial iron-comb structure such as described in [17]. However, these generators do not handle small oscillations well; because for complete flux reversal, the iron or magnets must move past several coil windings. This becomes very difficult when oscillation amplitudes are only a few millimeters because the size of the wires must be extremely small. The suspension and mass requirements to operate at higher frequencies in order to increase the power density are also not feasible [17]. Therefore, for this research, a moving iron type application is introduced with axial air gaps instead of radial air gaps; and the gas springs of the engine and air bearing provide all suspension requirements, thus reducing both losses and structural limitations. A more detailed description of the transducer designed for this research is provided in the following chapter. 1.5 Thesis Overview The basic operating fundamentals, history of thermoacoustic engines, and SOA have been provided in Chapter 1, as well as typical electroacoustic transduction methods in thermoacoustic applications. Chapter 2 provides a description of the thermoacoustic engine which is proposed for this research. Also in Chapter 2 is a description of the electroacoustic transducer in terms of the overall engine and how it is coupled to the proposed air bearing design. A detailed description of the modeling and design of the variable-reluctance generator, as well as the fabrication of the generator system is provided in Chapter 3. Experimental design and results for the static and dynamic tests of the variable-reluctance generator and subsequent analysis of the collected data is detailed in Chapter 4. Conclusions and suggestions for future work are provided in Chapter 5. 33 Chapter 2 Thermoacoustic Generator Description The project has been divided into three sub-components: the thermoacoustic engine, acoustic-to-electric transducer, and the bearing system for the transducer. A large portion of this thesis was dedicated to generating a viable concept for conversion from heat to electricity. Numerous design iterations took place in developing an understanding of these research areas. Ultimately, a standing wave engine with a linearly-acting, variable-reluctance generator is proposed while using self-pumping gas bearings for the linear alternator for both lifetime and improved suspension design. 2.1 Engine Design A schematic of the proposed thermoacoustic engine and electroacoustic transducer design is shown in Figure 2-1. The resonator uses a single piston to establish a mechanical resonance instead of an acoustic resonance. The single piston is used as the electroacoustic transducer converting the mechanical oscillation to electric power. In this configuration, the engine is simple and robust. The design is compact since it uses a mechanical resonance instead of the typical acoustic resonance. The rest of this section is dedicated to describing the design choices made for the thermoacoustic engine. 34 Heat In Heat In Heat Out Heat Out &I Bounce Volume Bounce Volume Pressure Vessel Wall Z Stack Stack I - I - Piston Stator W - - Flux Loop Windings - - - - - - - - - - Piston Rightmost Position Piston Leftmost Position Gas Bearing Surfaces It I Piston Steel Stator Fixed Inside Piston Minimum Air Gap Figure 2-1: This figure provides a diagram of the proposed thermoacoustic engine and acoustic-to-electric transducer. The acoustic wavelengths are significantly longer than any of the components with the piston acting as a center mass between two gas springs. The magnified views of the generator system shows the piston at the leftmost and rightmost extreme positions. 35 IC L I I C Figure 2-2: This figure gives the approximate representation of the thermoacoustic engine acting as a double Helmholtz resonator with a center mass and two bounce volumes. The engine components are significantly shorter than an acoustic wavelength, and therefore the lumped impedance model is applicable. This simplification does not model the impedance to flow from the heat exchangers and stack. Figure 2-3: This figure is taken from [22] where Swift describes a mechanically resonant standing wave engine using an axial air gap based electroacoustic transduction system. The engine concept presented for this research is essentially a double Helmholtz resonator where the pressure amplitudes across the piston are 180 degrees out of phase. Figure 2-2 gives the electrical analog for the resonator, where the capacitors are the gas springs, labeled "Bounce Volumes" in Figure 2-1, and the piston is the inductance. This mechanical resonance and transducer concept is similar to one proposed by Swift where the mass of the transducer is, as he says, "resonated away," and the highly efficient transduction from acoustic to electric power can occur through the nonlinear, inductance based transducer [22]. Figure 2-3 depicts Swift's concept of the mechanical resonant system with nonlinear transduction. Thus, the mass of the piston is necessary to the operation of the engine, and sets the resonant frequency with the gas springs. 36 Traveling Wave Design Constraint Contrary to what has just been described as this works ultimate engine design, the original engine concept was to use a traveling wave scheme such as those being designed for the SCORE project ([1] [5]) because of their potentially more efficient operation than standing wave engines as discussed in Section 1.3. These engines consist of a looped tube, traveling wave, regenerator and inline electroacoustic converter as previously shown in Figure 1-10. To decrease the size of the looped tube thermoacoustic engine and make it a portable system, either the tubing diameter or length must be decreased. Decreasing the diameter of the feedback tubing decreases the acoustic power for a given drive ratio and also increases the relative viscous losses due to the increased surface area to acoustic volume ratio. In short, both the power density and efficiency are significantly reduced by decreasing the size of the tubing to make a portable engine. Alternatively, decreasing the length of the tubing increases the frequency given the integral number of wavelengths required for a traveling wave around the engine loop. At higher operating frequencies, the mass of the electroacoustic transducer becomes an increasingly significant impedance, because the pressure difference required to move the piston with the same amplitude increases with the square of the frequency. Assuming a sinusoidal motion, the acoustic displacement (i), velocity (u), and acceleration (a) can be defined as (t) = Re[1ei" t ] (2.1) u(t) =- = Re[iw6iewt ] dt (2.2) (t(t) = 2 = e6(-21ei (2.3) where Re[] denotes the real portion of the complex number and 61 is the complex amplitude and phase of the displacement such that 1 = 6eeo 37 (2.4) where &, is the amplitude and # is the phase. This is the same notation as used in [24], and the "1" subscript will continue to represent the complex amplitude and phase of acoustic variables. In steady state, the pressure amplitude across the piston, Ap, is related to the displacement amplitude as 2 Ap A= w- mp~ AA (25 (2.5) where mp is the mass of the piston and A is the cross-sectional area of the piston. Typically, pressure and velocity variations across a lumped acoustic element are given in terms of an impedance defined as A(2.6) Z =zU 1 or Z= AU1 (2.7) depending on whether the component causes a change in pressure or volume flow rate. The mass of the electroacoustic transducer in a looped tube cannot be resonated away as in the standing wave type engine. The transducer is an acoustic impedance, which increases with the square of the frequency. This impedance causes a shift in phase between pressure and volume flow rate which must be corrected in order to maintain traveling wave phasing in the regenerator. This correction in the SCORE engine is done using a "stub" element following the transducer as shown in Figure 1-10, which is meant to tune the volume flow rate and pressure back into the appropriate phase [1]. This acoustic element creates a change in volume flow rate at constant pressure, where as the transducer changed the pressure at constant volume flow rate. However, the stub represents an additional loss mechanism proportional to the pressure times the change in volume flow rate. As the impedance of the electroacoustic transducer increases at higher frequency, the change in volume flow rates across the stub must also increase, with an additional increase in the losses of the system. Continuing to increase the frequency or mass of the piston causes the piston to 38 become the equivalent of a wall to the high frequency oscillations, and the resonance pattern in the tube becomes that of a standing wave. 2.1.1 DeltaEC Model To get a more accurate basis for the design of the engine and requirements for the power transducer, the Design Environment for Low-amplitude ThermoAcoustic Energy Conversion (DeltaEC) was used [311. This software, developed by Los Alamos National Laboratory, is specifically designed for numerical integration of the momentum and continuity equations for acoustics using predefined or user-defined acoustic segments and ensuring continuity of the segment boundaries. DeltaEC is useful for making design decisions, but provides only moderately accurate predictions of engine performance (usually within 10-20%). This is especially true for large amplitude oscillations, where non-linearities become more significant [23]. A DeltaEC schematic of the engine is provided in Figure 2-4. The engine has two sides that are symmetric about the piston power transducer. The components Hot". ?tsKHX WmdHX - "-m COM Dud Sectrscousic w T#rsdwv Figure 2-4: A schematic of the engine as designed in the Design for Low-amplitude ThermoAcoustic Energy Conversion (DeltaEC) program. Additional labels are included to identify components. The engine is symmetric around the center piston transducer. 39 Table 2.1: Thermoacousic engine parameters Component Description Variabl e Value Global Parameter Mean Pressure Frequency Gas Drive Ratio Engine Length Hot Duct Length Hot Duct Area HHX Temperature HHX Gas Area/Total Area Heat Transfer per Side HHX Length HHX Area Stack Gap Thickness Stack Length Stack Area -HX Temperature CHX Gas Area/Total Area CHX Length CHX Area Cold Duct Length Cold Duct Area Piston Mass PM 30 bar 250 Hz Helium 0.05 .26 m(10 in) 25 mm 2 4.56. 10-3 m Hot Heat Exchanger (HHX) Stack (Parallel Plates) Cold Heat Exchanger (CHX) Cold Duct Transducer Dr LHD AHD_ LHHX 668 K 0.67 194.12 W 5 mm AHHX 3.8. 10-3 TH # Hot Duct f Qrn y/o Latack AStack 4 LCHX ACHX LCD ACD_ mp m 2 0.14 mm 60 mm 2 3.8 -10-3 m 340 K 0.253 6 mm 2 3.8- 10-3 m 8 mm 2 4.56. 10-3 m 0.2 kg on each side of the transducer include a hot duct, which constitutes a large portion of the bounce volume, hot and cold heat exchangers (HHX and CHX respectively), and a stack. In the schematic, two stacks are shown on each side, which was done for modeling purposes so that the properties at the center of the stack could be readily identified and used for engine characterization and optimization. The "Begin" segment in the DeltaEC program initializes certain variables such as the mean pressure, and the "Hardend" segment allows the user to dictate that no acoustic power is flowing past that point (i.e. the acoustic impedance for both real and imaginary components are set to zero). The DeltaEC program is based on guesses, targets, and fixed engine parameters. For more information on the DeltaEC program see references [24] [31]. The DeltaEC code used for the design of this engine is provided in Appendix B. Design decisions were then made based on the DeltaEC simulations. The operating parameters of the model are provided in Table 2.1. These operating parameters are for a thermoacoustic engine that is 0.26 m (10 in) long and 0.0762 m (3 in) in diameter. This was determined to be a reasonably portable size. 40 These operating conditions were chosen after a number of optimizations were done using the DeltaEC program. The details of the engine design optimization are provided later in this section. The system, according to DeltaEC, behaves very similarly to the double-Helmholtz resonator depicted in Figure 2-2. This is clearly seen in Figure 2-5, where the signed amplitude of the pressure is shown and the magnitude of the volume flow rate. The pressure amplitudes are nearly 180 degrees out of phase on either side of the transducer, while the volume flow rate is a maximum at the transducer and nearly linearly decreases to zero at the ends of the gas springs. This is conceptually the same as the phasing described in the example standing wave engine shown in Figure 1-3, except that the piston mass has discretized the otherwise sinusoidal appearance of acoustic resonance in a tube closed at both ends. Also shown in Figure 2-5 are the acoustic and total power flows within the engine. Acoustic power is important because it gives the power available in the standing wave that can be extracted by the transducer element, and gives a reference for what dissipation elements are of greatest importance in the engine. The acoustic power is the time averaged power over an integral number of cycles given by the equation [24] 2 (x) Re[p1(x)ew t ]Re[Ui(x)ewt dt = 1 p1 ||U1 I coskPU 2 where 4pu (2.8) (2.9) is the phase angle between the complex pressure and volume flow rate. Therefore, examining the figure moving left to right, acoustic power is dissipated in the hot duct and HHX as the power drops below zero and then increases traveling through the stack portion of the engine. Acoustic power is then dissipated in the CHX and cold duct and then drops significantly across the VRG transducer element. The same acoustic power trend is mirrored on the other side of the thermoacoustic engine, but is negative given the 180 degree phase shift in pressure while the volume flow phase angle remains the same. Referencing Equation (2.9), the acoustic power is clearly dependent on the phase 41 Cold i I i Cold d I VRG| Duct |CHX|Stack|HHX Stack CHXXDuct- HHX Hot Duct -- - - --- -- I I S 200 Hot Duct -- Vessel Wall:l LPressure Pressure and Volume Flow Rate in Thermoacoustic Engine 150 100 50 0 '- -50 -100 -150 Pressure -200 Acoustic Power and Total Power C.) 200 U 0 4-J 0 Total Power 100 0 -100 -200 -300 -400 05. 15 1 2)5)r 2 . .4- 400 300 DistanceAlong Thermoacoustic Engine (m) . Figure 2-5: This figure provides the predicted pressure and volume flow rate profiles for the thermoacoustic engine, which indicate that the system behaves very similarly to a double Helmholtz resonantor given the 180 degree phase shift in pressure across the VRG. Pressure is given as the real component, Re[pieiwt}. Volume flow rate is given as the magnitude, I Ui Also shown is the acoustic and total power in the engine. The acoustic power only increases in the regenerator and decreases due to losses elsewhere. The total power is only changed by power entering or leaving the control volume of the thermoacoustic engine such as in the HHX, CHX, and VRG. 42 angle between the pressure and volume flow rate, #pu. For standing wave phasing, the phase angle between pressure and volume flow rate is nearly 90 degrees. The phasing cannot be exactly 90 degrees, because then no acoustic power would be generated in the stack. However, with a phase angle of nearly 90 degrees, the acoustic pressures or volume flow rates must be very large for meaningful acoustic power to be generated. This is another method of understanding the limitations of the standing wave engine compared to the traveling wave engine where acoustic pressure and volume flow rates are in phase and thus meaningful acoustic power can be generated at lower acoustic amplitudes. The total power flow in a thermoacoustic engine is also a time averaged value over an integral number of cycles [24], and is significant in thermoacoustic engines since, treating the engine as a control volume, it represents the net flow of power in and out of the engine. In more precise terms, the total power flow is the time averaged enthalpy flux and conduction through the solid/gas in acoustic segments where a temperature gradient is relevant. The total power term is then given by the equation ft 2 (x) = -pmRe[h101] - (Ak + Asolidksoid) dT 2 dx (2.10) where h1 is the enthalpy, C1 is the complex conjugate of the volume flow rate, A and Aolid are the cross-sectional areas in the stack of the gas and solid respectively, and k and k,olid are the thermal conductivity of the gas and solid respectively. For the model of the standing wave engine in this research, it was specified that energy could only be provided to or extracted from the engine through the heat exchangers or the electroacoustic transducer. For this reason, the total power must be constant everywhere else. In Figure 2-5, the total power is seen to be constant in the hot duct, increases in the HHX's, and decreases in both the CHX's and the electroacoustic transducer element where power was extracted from the engine. The efficiency of the engine, ratio of power extracted to the thermal power provided to the engine, is also apparent from this plot comparing the increase in total power across the HHX segment (approx. 200 Watts per side) to the decrease in total power (power generated) across 43 Figure 2-6: This figure shows the important geometric variables of the stack gas flow, where y, is the half width of the gas passage dimension and h" is the half width of the plate material. the VRG segement (approx. 50 Watts). The stack is the core of the thermoacoustic engine, where a portion of the total power is converted from thermal to acoustic power. Therefore, the parameters of the stack, including the stack placement and design/fabrication, significantly effect the power output and efficiency of the standing wave engine. For this engine, the stack was modeled as a series of parallel plates as shown in Figure 2-6, where h" represents the half thickness of the solid plates and yo is the half thickness of the gas passage gap. Typically, standing wave engines have stacks located at approximately A/20 from the pressure node located at either end of the half-wavelength resonator. This location has been determined to be the approximate required location to maximize the ratio of power output to viscous losses [24]. Shifting the stack closer to the velocity node Ij, and increases the pressure amplitudes, 1pi , decreases the acoustic displacements, I which effectively increases the critical temperature gradient as defined in Equation (1.4). Increasing the temperature gradient can result in increased conduction and diffusion losses and less power output. Conversely, shifting the stack closer to the velocity antinode results in larger volume flow rates and subsequently larger viscous losses. For this engine, which is designed to operate at 250Hz, the acoustic wavelength is between 4-6 meters using Helium as the working fluid. The range is given because the speed of sound, "a", has a significant dependence on the gas temperature, and 44 Effect of Stack Location on Efficiency TH 210 - Effect of Stack Location on 900 Max Efficiency Point7 S205 HD z 200 X 700 600 X195 5 10 15 20 190 5 10 15 20 Cold Duct Length (mm) Cold Duct Length (mm) (b) (a) Figure 2-7: The stack location is varied by changing the hot and cold duct lengths, but maintaining a constant total length. (a) The temperature significantly increases as the cold duct length increases. (b) Engine efficiency is also coupled to the placement of the stack to balance viscous losses with temperature gradient losses. the engine itself contains a large range of temperatures between 300-700 Kelvin. The acoustic wavelength would place the stack at approximately 0.25 meters from the end. However, because this system is being designed for portability, a 0.5 meter system would be ungainly. Therefore, the heat exchangers and stack were placed as close to the piston power transducer as possible. To verify that the stack should be placed near the piston and close to the A/20 value, a number of DeltaEC simulations were conducted to vary the length of the hot and cold ducts, while maintaining the overall length of the system. The piston displacement and transducer power output were also fixed for this optimization. Figure 2-7 shows the effect of moving the stack in the bounce volume while keeping the overall length constant, where the x-axis represents the length of the cold side duct. Thus, decreasing the length of the cold side duct is the same as shifting the relative position of the stack away from the velocity node. As expected, there is a local minimum in heat transferred to the engine (maximum in efficiency), and the hot side temperature continues to drop as the critical temperature decreases. The chosen design point was to place the stack at the location for maximum efficiency. Another way to ensure the proper acoustic impedance in the stack is to neck down the diameter of the engine at the location of the stack and heat exchangers. Figure 45 Geometry Optimization 222 Hot Duct Length 0.043m 221 0.04m 220 S219~ Z 0.036m 219 218 217 I Smallest Geometry Most Efficient Geometry 216 L9 15.2 11.4 7.6 3.8 0 Precent Reduction in HHX, Stack CHX Cross Sectional Area Figure 2-8: The diameter of the cross-section is necked down at the stack and heat exchangers to maintain the proper gas displacement amplitudes while minimizing engine size. The dimensions for maximum efficiency and minimum overall size are labeled for reference. 2-8 provides a summary of the geometry optimization, where the cross-sectional area for the stack and heat exchangers was varied, while also varying the length of the hot duct in order to determine an optimum geometry. In this simulation, the smaller the hot duct, the smaller the overall engine was, which is ideal for portability. The design points for minimum size and minimum HHX heat transfer (maximum efficiency) are labeled in the figure. Therefore, there is a balance that must be struck between minimizing the size of the engine and the engine efficiency. In addition to the placement and size of the stack, another factor in the efficiency and operation of the standing wave engine is the gap spacing in the stack, labeled "yo" in Figure 2-6. To maximize the engine efficiency, this stack spacing should be matched with the thermal penetration depth, which is inversely proportional to the square root of the frequency. Therefore, an optimal spacing was searched for in DeltaEC as shown in Figure 2-9, where the two factors considered were again the size of the engine and the hot side temperature in relation to the gap spacing. There is a clear efficiency optimum for the stack gap spacing at the point in part (a) of the figure where the HHX heat transfer to the engine is a minimum. This is a maximum efficiency because the power output from the transducer was 46 Stack and Geometry Optimization Stack and Geometry Optimization 800 240 Smallest Geometry/ 235 Hot Duct Length 0.05m Highest Frequency r230 22S 0.02m 750 Hot Duct Length O.03m O.04m O.05m 0.04m 2 2250.02m .550 220 521500 210 550 S205 -00 Most Efficient 0.1 0.12 0.14 0.16 Geometry .18 q 500 .2 0.1 Half Width of Stack Spacing (mm) .18 0.14 0.16 Half Width of Stack Spacing (mm) 0.12 .2 (b) (a) Figure 2-9: The efficiency of the standing wave engine is particularly sensitive to gap spacing in the stack. (a) The efficiency have local maximums where the heat transfer to the HHX is lowest. (b) As the hot duct length decreases and the frequency increases the hot side temperature increases significantly to compensate. held constant as the other variables were changed. The half gap stack dimension, y0 is strongly coupled to the efficiency, and therefore should be strictly controlled in the manufacturing process. However, manufacturability of the stack to exacting tolerances becomes increasingly challenging as the gap spacing decreases [24]. Further research is necessary for precisely controlling the plate spacing during fabrication to increase the actual efficiency of standing wave engines, especially for higher frequency applications. In part (b) of Figure 2-9, the hot side temperature is shown for different hot duct lengths. This again shows the effect of the critical temperature gradient as the acoustic displacements decrease in the stack and the pressure amplitudes increase as the stack is moved closer to the end of the hot duct. The length of the stack can also be used to optimize the engine. If the stack is lengthened, the critical temperature gradient remains approximately the same, but the actual overall temperature difference across the stack, set by TH and TC, must increase. If T0 is fixed, this indicates that the hot side temperature must increase as it did when moving the stack toward the the pressure antinode. However, with the increased length, more surface area is presented to the gas for heat transfer, and the 47 same acoustic power can be produced across a smaller temperature difference. This can increase the efficiency as discussed in Section 1.3, but also leads to additional viscous losses because the restriction to flow increases. This is another optimum that was performed in DeltaEC, but is not shown in a figure here. A number of iterations for each of these optimizations were performed in order to develop a system that would be compact, have a high power density with a reasonable drive ratio, and have a reasonable efficiency as predicted by DeltaEC. The final design is summarized in Table 2.1, but the full design (including dimensions, power, phase angles, etc.) is included in Appendix B. The power transducer used in the DeltaEC program was an "IESPEAKER" element placed in series between the two cold duct segments. This element is ideally suited for incorporating the moving coil type linear alternator. It became apparent from the DeltaEC model that the piston/gas displacement amplitudes would be quite small for a higher frequency/high mean pressure thermoacoustic engine. For this reason and others that will be outlined in Section 2.2, an alternative form of power transducer is investigated and constitutes the majority of the work for this thesis. A direct coupling of the transducer designed for this application and the DeltaEC software was not performed. The linear alternator segment of the model was therefore used to model as accurately as possible the mass of the piston and power output characteristics, without trying to incorporate the highly non-linear transducer that was actually developed during this project. Now that the transducer is built, a more accurate transducer DeltaEC segment could be defined using a user defined "RPN" segment, which is suggested for future work on this project, to more accurately predict the operating characteristics of the PTAG system. 2.2 Variable Reluctance Generator Design The DeltaEC models coupled with the project objectives led to very clear design requirements for the electroacoustic transducer. These requirements were low oscillating mass, small displacement amplitude, high power density, and efficient conver48 sion from mechanical oscillations to electric power. The system becomes necessarily higher frequency with small displacement amplitudes for three reasons: the internal gas volumes must be made as small as possible to make the system portable, the drive ratio (pi Il/Pm) must be less than 0.1 as discussed in Section 1.3, and the power output of the transducer is proportional to the operating frequency. To make the system portable, the "bounce volumes," as labeled in Figure 2-1, need to be as small as possible. Considering the thermoacoustics as merely gas springs for the moment, the effective spring constant for an isentropic compression is pm7IA , K= (2.11) where Ap is the cross-sectional area of the piston interfaced with the gas spring and V is the volume of the gas spring. Thus, the effective spring constant increases proportionally with the mean pressure and inversely proportional to the volume. The resonant frequency of the double Helmholtz gas spring-mass system is then simply . (2.12) F = Therefore, decreasing the gas spring bounce volume, as well as increasing the mean gas pressure, increases the resonant frequency of the system. Combining Equations (1.6), (2.5), (2.11), and (2.12), the drive ratio is Dr = 2 V 1j .Ap (2.13) In developing Equation (2.13), AP across the piston is assumed to be 2Ip1I, which is indicative of standing wave phasing, and I1i is taken as the gas displacement at the piston interface. Examining Equation (2.13) it becomes apparent that the drive ratio is not affected by the mean pressure, but only by the piston area, piston displacement, and bounce volume. Therefore, as the bounce volume decreases, either the piston area and/or piston displacement must also decrease. The force exerted by the gas on the piston decreases in proportion to the piston area, so it was determined 49 that the transducer should be designed to handle small oscillations to keep the drive ratio within the typically linear region. The piston mass does not affect the drive ratio, but it does lower the resonant frequency of the engine. Given the small oscillations of the transducer in order to couple with the thermoacoustics, higher frequency operation is necessary to extract meaningful electric power from the oscillations. This is because transducer power scales proportionally with frequency and approximately the square of the piston displacement amplitude. From this analysis, the conclusion was drawn that a portable thermoacoustic engine would require a small displacement, high frequency electroacoustic transducer. These requirements led to the design of a linearly-acting variable-reluctance generator (VRG) with an axial air gap. The axial air gap concept is different than typical linear alternator systems which rely on flux reversal and a constant radial air gap, "G" as shown in Figure 2-10 part (a). Shown in the figure, typical linear alternators use permanent magnets which travel past a full winding to completely reverse the flux passing through the winding. This provides efficient transduction using magnetic shear forces and a constant radial gap if the displacement amplitudes are large enough to completely reverse the flux. For smaller amplitudes, it is difficult to fabricate a a system where complete flux reversal is possible. This led to the alternative VRG geometry shown in Figure 2-10 part (b) where the flux does not reverse, but instead the axial air gap is changed, which causes a change in inductance. The VRG concept, at the most basic level, is to apply a current to the winding at large inductance and extract a larger current when the inductance is small. The increase in current is driven through the winding by the mechanical forces acting on the piston. This geometry is beneficial for small displacements because the inductance changes rapidly with an increase in air gap, "G". However, for the same reason, the VRG concept is not very good for large displacements because the magnetic forces drop with the square of the air gap length. Because there are two air gaps in this design for a given flux loop, the drop in force occurs even more rapidly. This leads to relatively poor coupling with the mechanical motion because the greatest velocity 50 Piston Leftmost Extreme 4 uo Piston Leftmost Extreme G winding Flux Loop IIJLH Piston Rightmost Extreme I G - sttow winding Pisston Rightmost Extreme G C n windi ing Flux Loop 06 G (b) (a) Figure 2-10: (a) Typical flux-reversing radial air gap linear alternator. (b) Small displacement axial air gap variable reluctance generator. occurs at a 90 degree phase shift from the piston extreme locations and minimum air gaps where the largest magnetic forces are generated. However, the magnetic normal forces can be an order of magnitude larger than the shear forces as described in Section 3.2. Therefore, for small oscillations, the large magnetic forces can be maintained efficiently through the length of the oscillation. The diagram shown in Figure 2-10 is one half of the VRG concept, which actually has two connected sides so that the transducer extracts power during both halves of the piston sinusoidal motion. A more detailed drawing of the VRG system shown in Figure 2-1 is provided in Figure 2-11. It is relevant to note that the generator can only experience attractive forces between the stator and piston components. Therefore, the mechanical action is always to push from one side of the piston to increase the air gap on the opposing side. The 51 Piston Piston Steel Stator Windings Stator Fixed Inside Piston Figure 2-11: The VRG component from Figure 2-1 is shown again here for reference. The sinusoidal pressure acting on the piston steel surfaces are 180 degrees out of phase as the piston oscillates around the fixed internal stator changing the air gap, "G", on both sides. force is transmitted through the piston support structure, labeled "Piston" in Figure 2-11, which holds the two sides of the piston together. Additionally, the stator is fixed internally to the piston so that the piston oscillates around the stator. The modeling, construction, and experimentation of the VRG are described in much further detail in Chapter 3. This description is provided to give an understanding of the proposed engine design and to give a basis of understanding of what is needed for the gas bearing, which is the topic of the following section. 2.3 Gas Bearing Design The easiest way to describe the operation of a gas bearing is with a voltage divider analogy as shown in Figure 2-12. In the figure, a pressure reservoir at pressure PHigh supplies gas through an orifice to the gap between the piston and the cylinder. The gas flows through this gap to the edge of the piston to a low pressure reservoir ) at PLow. The fluid flow resistances provided by the orifice (R1 ) and the gap (R 2 constitute a "voltage" divider. The pressure at the orifice-gap interface, Pmeaing, is 52 Piston wall Orifice 0 R1 PHigh R2 RR Figure 2-12: This figure depicts the voltage divider analogy for gas bearing design. This was the basis of analysis for evaluating self-pumping gas bearing concepts. easily calculated using the analogy of the voltage divider, PBearin ge 2 R2a+nR1 (PHigh - PLow) + PLow (2.14) The value of PBearing is dependent on the gap dimension "g". Should the piston move towards the cylinder wall, the resistance to flow in the gap, R 2 , increases. Since the orifice flow resistance, R 1 , and reservoir pressures, PHigh and PLO, remain unchanged, Equation (2.14) requires that PBearing increases. The reverse is also true. Should the piston move away from the cylinder wall, the gap resistance decreases. This drop in resistance leads to a subsequent drop in bearing pressure in accordance with the voltage divider. In a piston configuration, the reduction in gap on one surface of the cylinder wall is accompanied by an increase in gap on the opposing surface. Since the overall pressure in the gap has increased on the side with a smaller gap and decreased on the side with a larger gap, the net pressure force on the piston tends to restore the piston back to the original centered position. This restoring effect is most responsive when the orifice resistance, R 1 , and gap resistance, R 2 , are approximately equal in the centered piston position. To approximate the flow resistances for the purpose of modeling specific bearing 53 geometries, circular Poiseuille and plane Poiseuille flow were used for the orifice and gap respectively. The plane Poiseuille resistance is approximated by the equation (2.15) R = 12 1L3 wg where p, is the viscosity of the gas and w and L are the width and length of the channel respectively. The circular Poiseuille flow resistance was approximated by the equation R = _128iLc rD4 (2.16) 7rD4 where D and L, are the diameter and length of the orifice respectively. Typical gas bearing systems are fed directly by an external pressure supply or compressor to operate continuously. For a portable power system, the aerostatic bearing cannot be fed by an auxiliary system. Therefore, a number of methods were considered to produce self-pumping gas bearings, which relied on the oscillating pressures to feed high and low pressure reservoirs or plenums. The design concepts and design chosen for further development are described in the following sections. 2.3.1 Design Iterations The idea of self-pumping gas bearings was initially considered after reading through the Swift and Backhaus paper on gas diodes [25]. The gas diodes have been used both in jet pumps for preventing Gideon streaming and for self-circulating heat exchangers. The concept is to use non-symmetric gas passages to change the resistance to flow between the forward and reverse flow directions. This is done by effectively changing the minor loss coefficients, K+ and K_, in the forward and reverse directions respectively. The minor loss coefficients express the irreversible turbulent pressure drop across an impedance element as defined by 12 AP = K-pu2 2 54 (2.17) *Diodes treated as nonlinear resistors Po~sc WW ating Piston Motion PISTI Piston Rs* PLow PHigh Bearing Surface Stator Bearing Surfaces Figure 2-13: This figure depicts the gas diode based bearing design for both the nonlinear end effects and piezoelectric check valve concepts. The resistance R 2 constitutes the orifice resistance, and the resistances R 3 and R 4 in parallel constitute the second voltage divider resistance depending on whether POcillating is high or low. where K is equal to K+ when the flow velocity, u, is positive, and K is equal to K_ when the velocity is negative. In oscillatory flow, this leads to a net DC flow in one direction if K+ is not equal to K_ [25]. This DC flow component is exactly what is necessary to produce a gas bearing system. Common to all the designs presented here are internal pressure reservoirs to act as accumulators to rectify the AC pressure oscillations to DC flow. In the electric circuit analogy, these reservoirs act as "capacitors". One inherent reservoir in the design presented for the VRG is the volume inside the piston, which also contains the stator. The second reservoir presented in these designs is a secondary chamber fabricated into the cylinder wall. A schematic of a bearing concept using the gas diodes similar to the Swift gas diodes is shown in Figure 2-13. This figure shows a section of the piston, where the piston has a bearing surface on which the pressure, PBearing acts, and changes with respect to piston motion off its center-line axis. The bearing surface is a ring around the circumference of the piston, and although only one is shown in the inset of the figure, a second bearing surface is present on the lower half of the piston as well. In 55 Pressure Regions as a Function of Time Pressure Regions as a Function of Time 3.15 3.1v 75%I61mm - 2.__5 - - M High PinmCtkn" P_ [ 1 26 2.9 2.86 - s nn """ Hih t P,41h, 9 2.9 0 D002 004 0.006 0.008 0.01 0.02 299 0 0 0 6 Time (s) Time (s) (a) (b) 0 0.01 1 0012 Figure 2-14: (a)This figure shows the oscillating pressures achieved with a piezoelectric gas diode and a passive end effect diode on the piston. Pressures are the same as indicated in Figure 2-13. Shown in (b) is the variation in bearing pressure as the piston moves off its center axis to increase or decrease the radial gap between the piston and the chamber wall. the figure, the resistances R1 and R5 use asymmetric channels to produce a net DC flow, which tends to "charge" or "discharge" the high and low pressure plenums. With the high and low pressure plenums established, the orifice resistance, R 2 , and bearing gap resistance, R 4 , constitute the voltage divider necessary for the operation of a gas bearing. Thus, a decrease in the bearing gap causes the resistance R 4 to increase resulting in an increase in bearing pressure, PBearing. The reverse is also true, and the piston tends to remain centered due to these changes in bearing pressure. The connection of the resistance, R 4 , to the reservoir inside the piston labeled PL,, is accomplished by the same ports in the piston through which the stator is connected externally to the pressure vessel wall. Unfortunately, for the passive DC flow design based on asymmetric end effects, the reverse bias on the diode is quite poor. For example, the minor loss coefficient in the forward and reverse directions, K+ and K_, can be changed by a factor of approximately 2-3 depending on the chosen geometry. For large pressure oscillations, this leads to a significant breathing effect for gas flowing in and out of the plenums. This was calculated to dissipate several watts of power based on the pressure drops 56 and volume flow rates. For this reason, it was proposed to introduce a piezoelectric element to further increase the difference in minor loss factors between forward and reverse flow. This concept holds merit, but the strain values for piezoelectric material are on the order of 500pm/V. This indicates that extremely small holes would be necessary for the piezoelectric strain to cause a change in the gas flow resistance. These small holes would significantly decrease the maximum pressures obtainable in the plenums. Additionally, the gas diode located on the piston as shown in Figure 2-13 could not easily be of the piezoelectric type, because flexible wires would need to be connected to the piston and represent a failure mechanism. The piezoelectric concept still appears to be a valid solution if the piezoelectric material can change the minor loss coefficient by a factor of 10 and a passive end effect diode is used on the piston. Figure 2-14 shows what plenum pressures and bearing pressures that could be obtained with 30 bar mean pressures and 1.5 bar amplitude oscillations if the piezoelectric material changes the minor loss coefficient by a factor of 10, and the engineer is willing to accept a few watts of dissipation to make the bearing operate. However, these limitations and losses coupled with the additional complexity, control and necessity for a high voltage across the piezoelectric material led this research toward other passive gas diode concepts. 2.3.2 Final Gas Bearing Concept The gas bearing design chosen, which has become the topic of another master's thesis in the Cryogenics Laboratory at MIT, uses the oscillating motion of the piston to act as the check valve for pressurizing the gas bearing system. A literature review of this concept yielded little in the form of papers, but did reveal a few patents related to this concept [7]. Typically these bearing systems have been suggested for free piston Stirling engines such as those developed by NASA for solar electric power [8]. For this research, two concepts for the sliding piston check valve were introduced. The first was ultimately a poor design, but led to the conclusion that both a high pressure and low pressure plenum were necessary for continuous bearing operation. Figure 2-15 depicts the original sliding piston check valve bearing system. 57 Bounce Volume 1 Piston ROscillatingl Motion R R5 Piston Port R4 High Pressu re PHigh Plenum *Stator Not S iown POscilIating2 (1810 Shift from Poscating) Bounce Volume 2 Figure 2-15: This figure shows the piston and cylinder wall where the internal portion of the piston is used as a high pressure plenum. The resistance R 2 provides the constant resistance pressure drop in the voltage divider analogy. Resistances R 3 and R5 increases or decreases as the piston moves closer or farther away. The resistance R 4 decreases to a near zero value as the piston moves up. At this point, the value of POscillating is at a maximum and gas flows from bounce volume 1 to the high pressure plenum, charging it to high pressure. This bearing system tends to charge the high pressure plenum, represented as a capacitor in Figure 2-15, because the resistance labeled R 4 decreases to a near zero value as the piston travels upward. In the piston's extreme upward position, the pressure, Posciiating1, is at a maximum and charges the plenum through the low resistance path labeled R 1 . The "voltage divider" for the journal bearing operation is provided by the constant resistance R 2 in the piston wall and the two resistances R 3 and R5 , which vary with radial piston displacement. This bearing system works for nearly the full piston cycle, but breaks down when the oscillating pressure, Posciiiatingi, is greater than the high pressure plenum, which occurs at the extreme positions of the piston. A schematic of the flow which causes bearing failure is shown in Figure 2-16 and plots the bearing pressures at various radial piston displacements. When the 58 Bearing Pressure 6 POscillatingl > PBearing1>PHigh x10 . When 3.14 Oscillatingi PISTON / '7 B 3.13- / R3 3.12 Fhat causes bearing to fail CL3.11 /Bearing Failure 3.1 Baing1 3.09F 33% Clearance Decrease Aligned n Clearance Increase -33% 3.08 . 0 ~2 3 4 5 6 7 8 9 10 Time (ms) (b) (a) Figure 2-16: (a)This figure shows a schematic of the flow which causes the bearing to breakdown. A flow reversal in R 2 reverses the voltage divider and makes the bearing unstable. (b) Shows the variation in bearing pressure as the piston moves off its center axis. The bearing becomes unstable as shown in the figure when an increase in the clearance gap increases the bearing pressure. This effectively forces the piston to touch down. oscillating pressure is larger than the bearing pressure, the flow is able to reverse and flow into the high pressure plenum through resistance R 2 . This effectively reverses the voltage divider and causes the bearing pressure to increase as the clearance gap increases causing unstable bearing operation. From this initial sliding bearing design, it was determined that a second low pressure plenum was necessary to ensure continuous bearing operation. A diagram of this bearing scheme is illustrated in Figure 2-17, where the high pressure plenum is still internal to the piston, but a second low pressure plenum is introduced into the wall surrounding the piston. The high pressure and low pressure plenums are both controlled by selectively opening and closing channels when the bounce volume pressure is at a maximum or minimum pressure. The relative phasing of the charging and discharging of the plenums is shown in Figure 2-18. In the figure, when the piston is in its lowest position as shown on the page, the pressure, POscillting,is at a minimum and PLO, drops to match this pressure. As the piston moves back up past 59 POscillatingl POscillatingl PISTON Rc R3 - PLow In Bearing3 . R6 aril10 bR2 R ( 2High PHigh (a) (b) Figure 2-17: This figure shows: (a) the proposed gas bearing system with both high and low pressure plenums (b) the variable references for the electric circuit analogy used in the analysis of this bearing system. "Capacitors" Charging "Capacitors" Discharging POscOlatngl POscillatingl ..*Resistance goes to zero Ptow R3 PBearing PISTON PBearing1 Beancg2 = goth *Resistance goes to zero -- PHigh T (a) High (b) Figure 2-18: This figure provides the relative piston position for charging and discharging of the plenums. Both plenums charge and discharge at the same time. Charging of the high pressure plenum occurs when Poscillatingl is high and the resistance R4 goes to zero. Discharging of the low pressure plenum occurs when Poscillatingl is low and the resistance R3 goes to zero. 60 Table 2.2: Gas Bearing Operating Parameters Component Description Variable Value Global Parameter Mean Pressure Frequency Gas Density Viscosity Temperature Radial Clearance Piston Displacement_ Volume Pm 30 bar 250 Hz Helium 4.6 kg/M 3 2. 10-5 Pas 300 K 12.7 /im 5 mm 3 10-5 M Plenum (PHigh) Plenum (PL,,,) Resistance (R4) ) Resistance (R 2 ) Resistance (Re) Re-si stanJe _(R 3 ) ) ) Resistance (R 4 Resistance (R 5 Resistance (R 6 Volume Length Diameter Length Diameter f p AI g Li D1 _ Length Effective Width (Same for R 4 -R6) Length Length Length _ _ 10_6 m 4 mm .8 mm 1mm 80 pm 3 L3 w 0 - 4.9 mm L4 0 L_6 _ _ _ 1 - 5.5 mm 1_- 5.5 mm 3 mm - 4.9 mm the low pressure port, the resistance R 3 increases rapidly and the flow rate into the port is less than in the discharge configuration. The flow path through the RI? resistance from the high pressure plenum to the low pressure plenum provides a constant bearing surface and ensures no flow reversal through the R 2 resistor, which would cause the bearing to fail. Therefore, a continuous gas bearing is formed on both the upper and lower bearing surfaces of the piston. A number of assumptions were made for the analysis of this bearing system. The plane Couette flow was not included even though the oscillating piston velocity would cause some flow due to the no slip boundary conditions. Additionally, the gas compressibility was excluded from the analysis because the relative changes in pressure were small compared to the mean system pressure. Table 2.2 lists the pertinent dimensions for the final bearing system shown in Figure 2-17. Gas bearings typically require tight tolerances and honing of surfaces. The radial gap dimension given in Table 2.2 was based on discussions with Professional Instruments. Typical radial bearing clearances were said to be within 12.7 pm and 25.4 pm (0.5-1 thousands of an inch) on the diameter for the 0.0762 m (3 in) diameter piston. 61 Restoring Force Gas Bearing Spring Constant 45 40 30 - 35 frs= 900-1100 Hz 30 020-~ 25 U020 10- - ~30 ift - 50% Shift 70%SN -4-Low Force -0 Hg oc -High Force Linear (Low Force) 10 Shi -90% - 5 5- 1 0 0 0 0.5 1 1.5 2 Time (s) 2.5 3 3.5 X1- 2 4 3 Linear (High Force) 5 6 7 Piston Displacement From Center (pm) 4 (b) (a) Figure 2-19: (a) The restoring force acting on the piston for different radial displacements is shown over one cycle. The restoring force changes over the cycle due to the fluctuating bounce volume pressures, resistance lengths and plenum pressures. (b) The restoring force effective spring constant is calculated for both a best and worst case scenario based on the restoring forces shown in part (a). The spring constant leads to resonant frequencies significantly higher than the proposed 250 Hz piston operating frequency so lateral resonances should not be an issue. The pressure was assumed to vary linearly along the piston face. The pressure profile is dependent on the bounce volume pressure and the pressure PL"". Thus, at each point in time, the three point pressures PBearingl, PBearing2, PBearing3 were determined and integrated along the length between the points to determine an approximate restoring force per unit width for the piston. A 3-dimensional model was not completed as part of this analysis, and therefore, the force per unit width was multiplied by the radius to get an estimate of the piston restoring force. This restoring force could be determined as a function of piston radial displacement from its centered position. The restoring force as a function of time at specific radial displacements and the effective spring constants obtained from these radial displacements is presented Figure 2-19. As shown in part (a) of the figure, the restoring force from the bearing can reach 20 - 30 N, which is more than enough for the 0.2 kg piston. Additionally, the resonant frequency of the bearing system must necessarily be much larger than the oscillation frequency of the piston. Shown in part (b) of the figure, 62 the resonant frequency of the bearing spring constant is approximately 5 times the oscillating frequency of the piston. Therefore, harmonics in the lateral motion of the piston should not be an issue with the design. It is proposed, but not sufficiently analyzed, that this same bearing system would act as a thrust bearing. This effect arises from the shared high pressure capacitor between the two gas springs. Should the piston drift off axial center, the high pressure plenum will be open to the gas spring longer, and the mean resistance to the low pressure plenum will me less than on the other side. This trust bearing application in addition to the analysis of bearing start-up and lift-off should be considered in future work. 2.4 Chapter Summary A full thermoacoustic engine and electroacoustic transduction method has been proposed in this chapter. The thermoacoustic engine has been modeled and to an extent optimized for the PTAG system. The DeltaEC program predicts an efficiency of approximately 12.5% from thermal to electric power. The electric power is generated by a variable-reluctance generator, which will be discussed in length in the following chapters. Finally, a gas bearing system has been proposed that can be coupled with the oscillatory motion of the power transducer piston in order to create a self-pumping gas bearing system. 63 Chapter 3 Design of Variable Reluctance Generator The Variable Reluctance Generator (VRG) design as briefly introduced in Section 2.2 was chosen for its potential to efficiently convert small-amplitude mechanical oscillations to electric power. This chapter presents the design, model, and fabrication of the electroacoustic transducer necessary for converting the thermoacoustic oscillations into electric power. 3.1 Design The stator for the electric transducer is shown in Figure 3-1 part (a). The stator consists of four laminated steel wedges as shown in part (b) of the figure. Four laminated piston wedges align with each of the stator wedges on each side of the stator as shown in Figure 3-2, giving a total of eight piston wedges and four stator wedges per VRG system. The two sets of piston laminations (one set on each side of the stator) complete the magnetic flux loops established by the two coils set into the stator. One set of piston laminations and the corresponding coil constitute one phase of the two phase generator system. During the operation of the VRG system, the two sets of piston laminations are driven sinusoidal closer and then farther away from the stator. This motion changes 64 (a) (b) Figure 3-1: (a) The stator design is shown, which consists of (b) the four laminated steel wedges. Stator Lamination - Phase 1 Piston Laminations Center Air Gap Surface Phase 1 Flux Loop Outer Air Gap Surface Phase 2 Piston Laminations - Phase 1 Coil - Phase 2 Coil Figure 3-2: The VRG magnetic core is shown consisting of 4 laminated steel stator wedges and 8 laminated steel piston wedges. The two sets of piston wedges align on either side of the stator to complete the magnetic flux loops. The VRG consists of two phases, which generate power by changing the inductance of the flux loop by moving the piston closer and farther away from the stator in a sinusoidal motion. 65 Ip Piston Laminations 1W hwPhase ICON _A, Winding Phase 2 CoHl Section AA' (a) (b) Figure 3-3: This figure depicts (a) the top down view of the stator with the associated geometry variables used for modeling and optimization. The piston is not shown because it exactly overlays the stator. (b) The side view of one of the four wedge stator pieces is shown and the corresponding piston wedge piece. the distance between the piston lamination face and the air gap surfaces labeled "Center Air Gap Surface" and "Outer Air Gap Surface" in Figure 3-2. The change in the length of this air gap changes the inductance of the VRG in a predictable way such that electric power can be extracted from the mechanical motion. The model of this change in inductance and the subsequent electric power generation are presented in Section 3.2. The dimensions of the system effect the inductance power output capability, and efficiency of the VRG system. The important dimensions of the VRG system used for modeling the transducer are shown in Figure 3-3. These dimensions were also used for optimization of the VRG system, which is presented in Section 3.4. For reference, the final dimension values selected after optimization are presented in Table 3.1. To interface with the thermoacoustic system, a continuous surface must be presented to the bounce volume as shown previously in Figure 2-1. Additionally, the two sets of piston laminations must be connected so that the gas pressures in one bounce 66 Table 3.1: VRG final design dimensions Component Variable Value Stator l4 w z hw 7.62 mm 22.22 mm 7.62 mm 11.81 mm 2.18 mm 15.88 mm -h 1P hp w 6.35 mm 35.56 mm 3.81 mm 11.81 mm 1W 10 Piston volume force the piston laminations on the opposite side away from the stator as discussed in Section 2.2. Therefore, the piston laminated steel components are inset into a piston cap, and the piston cap is fixed to a honed aluminum piston as shown in Figure 3-4 (a). The piston cap provides the pressure bearing surface interfacing with the bounce volumes, and the piston provides the connection between the laminated piston sets, as well as the required gas bearing surfaces as described in Section 2.3. When the full system is put together, as shown in Figure 3-4 parts (b) and (c), the stator sits internally to the piston and is held fixed by a screw or pin connected to the four "Stator Brackets" as labeled in the figure. These brackets protrude through the piston through the four ports in the piston. The piston mid-section between the gas bearing surfaces is relieved to allow machining of the ports without concern for the micron type tolerances required on the gas bearing surface. The VRG system may then be fixed between the two bounce volumes such that the piston oscillates with the stator enclosed within the piston. The following section presents the model used to predict the performance of the VRG when the piston is forced to mechanically oscillate in resonance with the thermoacoustic system. 67 (a) Piston Cap Phase 1 Piston Laminations Stator Laminations Stator Bracket Coil Phase 2 Piston Laminations (b) (c) Figure 3-4: (a) An exploded view of the piston, piston cap, and laminated piston steel components is shown. The piston assembly is designed in three parts so the stator can be fixed internally to the piston. (b) A radial cross-section of the full VRG system is shown where the stator is fixed externally inside the piston by the stator brackets. This figure includes all of the structural non-magnetic components that were not included in Figure 3-2, which leads to (c) the final VRG design. Not shown is the cylinder in which the VRG is fixed. 68 3.2 VRG Model 3.2.1 Linear Model Figure 3-5 is a radial cross-section of one phase of the proposed electroacoustic converter. It consists of laminated steel stator and piston elements and a copper coil as described in Section 3.1. Although saturation effects in a high power density device will be important, great insight and guidance towards a final design can be obtained by a simple linear model of the magnetic circuit used in this device. A magnetomotive force, F, equivalent to voltage in a circuit analogy, is generated by the net current flowing inside the magnetic circuit loop or F = NI (3.1) where N is the number of turns penetrating the loop and I is the current in each of those turns. The magnetic reluctance of a material in the loop is Rm = (3.2) [LA where 1 is the length of the magnetic flux path, A is the cross-sectional area of the material through which the flux passes, and p is the magnetic permeability. Often, 15 vVAv __ Piston Steel =Air Gap Air Gap Winding ^w~w^4stator Figure 3-5: This figure shows the magnetic circuit analogy used for the initial analysis of the VRG. Only half of the VRG is shown and is a 90 degree rotation of Figure 2-11. 69 the magnetic permeability of a material is given as a relative permeability such that A = pro where bo is the permeability of free space. The magnetic flux 0 is the equivalent of current in a circuit analogy and is given as .F m= Z Rm (3.3) where E R, is the sum of magnetic reluctances in a series circuit of reluctances. For linear magnetic materials, the flux linkage is linearly dependent on the current in the windings or A = LI (3.4) where L, the proportionality constant, is the inductance of the device. The voltage that appears across the terminals, V, of a device with no internal resistance depends only on the flux linkage as (3.5) V = -dt For the linear model of the electroacoustic transducer, the inductance can be described with a fixed steel reluctance and a variable air gap reluctance such that L= N Rsteei + Rgap (3.6) where Rgteel is the reluctance of the steel and Rgap is the sum of both the inner and outer air gap reluctances. Therefore, the inductance is only a function of the air gap length, G, which is given by the equation for sinusoidal piston motion as G= 2 cos(0) + 2 1+ Gmin (3.7) where 111 is the amplitude of the piston displacement,9 is the cyclic angular displacement, and Gmin is the air gap length when the piston laminations are at their closest position to the stator air gap surface, also referred to as the minimum air gap. To calculate the constant reluctance of the steel and the variable reluctance of the air gaps, the geometry of the generator must be known. Applying the notation 70 defined in Figure 3-3 and making the simplifying assumption that both the inner and outer air gaps have the same area, the following geometric relations can be made: A = 4(l - z 2 ) (3.8) w = 2(l4 - z) (3.9) 10 Ag = 4w (3.10) The reluctance of the steel and air gaps are then approximated by Rsteei = 2h' 1p 1p pAg phew phpw Rgap 2G - poAg (3.11) (3.12) where Rgap is the sum of both the inner and outer air gap reluctances. With the reluctances set by the physical dimensions of the generator, the inductance can be plotted as either a function of air gap length (piston displacement), G, or (using Equation (3.7)) the angular displacement, 0, as shown in Figure 3-6 and Figure 3-7 respectively. Since the energy stored in the magnetic circuit is A = 1 L12 2 (3.13) the force at constant current can be determined F dA =I1 2dL dG 2 dG (-4 Ultimately, the most important factor for the engine design is the power output capability. This is determined by the area of the flux linkage-current profile. The idealized current profile for the linear model can be determine by applying equations 71 Inductance vs Air Gap Length 5- 4- 3 -- 2 1 2 Air Gap Length (m) 6 x 10- Figure 3-6: This figure depicts the inductance for one phase of the generator as a function of the air gap length. This is based on the final design with 148 turns. (3.4), (3.5), and (3.7) to get 1(6) = VdO j""L(9) (3.15) where Oon is the phase angle at which the voltage is first applied across the winding. This leads to current profiles that qualitatively look like Figure 3-7 part (b), and fluxlinkage-current profiles that look like Figure 3-8, where the shaded area represents the work output per cycle converted from mechanical to electrical energy. Saturation and eddy current losses shift the flux-linkage-current profiles away from the ideal ones shown in Figure 3-8. These effects, discussed in Section 3.2.2, reduce the power output from the power outputs predicted by the simple linear model. Rotational Stability An additional benefit of the wedge design, as presented in this thesis, is the rotational stability provided by the cross shaped laminated steel components. Because the stator is held internally to the piston, if the piston rotates about its center axis, rubbing between the piston and stator brackets could occur. However, in this design, should 72 Cyclic Inductance Generating Motoring 5 4 C 3 2 0 1 5 4 3 2 6 Theta (rad) (a) 1 (0) Current: Motoring Oof 1(0) Current: Regeneration Oaff eon 0 0 (b) Figure 3-7: This figure depicts (a) the cyclic inductance with the assumed sinusoidal motion. Also shown are the regions for motoring versus generation. This is based on the final design with 148 turns. Provided in (b) are qualitative waveforms for the system operating as a motor and a generator. Adapted from [30]. 73 A L~e)I I Figure 3-8: This figure depicts ideal flux linkage-current profiles for operating the VRG. Current is provided at the maximum inductance and extracted at the minimum. Figure adapted from [30]. the steel components become unaligned, a net restoring torque will be generated to minimize the reluctance of the flux path given by 1 2 dL 2 dOr T- (3.16) where 0, describes the rotation of the piston about its center-line axis. Neglecting the reluctance of the steel and any fringing or leakage, the inductance is simply L = N 2 poA 2G (3.17) and the area as a function of 0 , as approximately A = 4(w - Orlp)lo. (3.18) Therefore, the restoring torque is approximately T- __I 2 N 21 ull * G * N 74 (3.19) The restoring torque is highly dependent on the gap displacement and current, but reasonable torque values were calculated to be between 0.1 - 0.9 N -m. Given that there should be no external forces causing rotation, this was determined to be more than enough to keep the piston aligned for the rotational degree of freedom. Therefore, the stator support structure could be passed through the ports in the piston assembly without concern of wear due to rubbing between a misaligned piston and stator bracket. 3.2.2 Nonlinear Saturation Model Saturation effects can be accounted for by using a piecewise linear model, an approach similar to that taken in [30]. In this piecewise linear model, the same linear inductance methods are assumed until the steel becomes saturated. Saturation occurs-for fields of 1.5 to 2.4 Tesla, depending on the material. In this study, a value of 1.7 Tesla was used for the initial generator design. After saturation, the saturated incremental inductance, L., is assumed to be constant and a factor of 1000 greater than the permeability of free space. This was to approximate the knee seen in the B-H curves for the saturation of steels. The steel saturates at the same flux linkage given by A = NBmin[Astee] (3.20) where min[Asteel] corresponds to the smallest cross-sectional area of steel along the flux path. This occurs in the piston steel because of the emphasis to minimize the overall piston mass. With this model, the flux linkage is described by the piecewise equation: A 0 < I < i(3.21) L(0)I(0) L(0)Is + Ls(I() - Is) Is < I. This leads to the characteristic flux linkage-current plots shown in Figure 3-9. These plots are for various piston displacements and are significantly different than the plots obtained by Vallese for his rotary variable reluctance motor (VRM) because the 75 Flux Linkage- Current Plot X 0.5unsaturated region Saturated Region .045- M 035 Unsaturated 0.5mm Region 0)" -. saturnted region - .04 I M 03- 025 -- X 4mm 's La M Lmax u'.02- 5.5mm .015.01-L .005L 095 10 15 Current (A) (b) (a) Figure 3-9: This figure depicts (a) the modeled flux-linkage-current plots at various piston displacements. This is in contrast to (b) which shows the model done by Vallese for a rotary motor, where the steel saturates at constant current [30]. steel becomes saturated at a constant current, rather than at a constant flux linkage as is the case for the linearly-acting VRG [30]. This is because the air gap area is fixed for the linearly-acting VRG machine where it is varying for the rotary type motor/generator. The force can then be calculated from the coenergy expression d F = -O i A(I, O)dI. (3.22) Having determined the flux-linkage-current characteristics of the generator, the force takes the same piecewise form dependent on whether the generator is saturated or not. The force is given by F= f I tI- I)Lcf 0 1 I 1I. (3.23) From these equations, the magnetic force on the piston can be calculated as a function 76 Force vs Displacement at Constant Currents -50-- 0 -100. -l=1A -1=3A -150 -1=5A -- -1=7A 1=9A -1=11A -=13A -200 -1=1 5A -250- -300 Air Gap Length (m) X 10- Figure 3-10: This figure depicts the the magnetic attractive force between the piston and stator at the various piston displacements with lines of constant current. of the position and current. Figure 3-10 shows what the static forces on the piston should look like along its range of motion at constant current profiles. These forces were used in subsequent design of components and experimental apparatuses to ensure proper mechanical strengths, deflections, and deformations. As seen in Figure 3-10, the force at saturation is constant (approximately 275 Newtons). This is true when neglecting fringing. A flux tube analysis was performed to determine the relevance of fringing for this generator, which will be discussed further in Section 3.2.4. While the linear model also assumed negligible wire resistance as in Equation (3.5), the actual voltage applied across the winding at a specific phase angle contributes to both changing the flux linkage and driving current through the winding resistance. For sinusoidal motion, this is given as dA Vp = I(6)R + WO (3.24) where the time derivative of flux linkage has been converted to d\ = dA dO d d dt 77 (3.25) Flux Linkage-Current Cycle Profiles Saturated and Unsaturated Current Waveforms zt. U.UD) Saturation Current Peak 18- 0.045- 16- 0.04 14 Saturation Current Peak 0.035 Min Air Gap . 120 10 .0 0.025- 0 cm x 8- 6 0.02 0.02- 0.015- 4- 0.01 - 2- 2 -Saturated Unsaturated .- 0.005 -- 0 0 1 2 3 4 Phase Angle (rad) 5 6 S 7 2 (a) 4 Current (A) 6 8 10 (b) Figure 3-11: Shown in the figure is (a) potential current waveforms for one phase of the generator. Examples of current waveforms for both saturated and unsaturated operation are shown, and (b) the flux-linkage-current cycle profiles for two potential excitation schemes (the same as in part (a)). Shown in the background are the plots of flux linkage and current at fixed axial piston displacements. and the time derivative of 6 is simply the angular frequency, w. Solving Equation (3.24) for flux linkage gives A(0) 1 rO [V - I(9)R]d6. (3.26) To determine the current profile, an initial current was assumed and the flux linkage was calculated. The current was then calculated using the piecewise equation -M () 0< I< i (3.27) L() I(M)-L()1 8 + is is <I. This process was iterated to determine the current profile accounting for saturation and the effect of the winding resistance. Figure 3-11 part (a) gives examples of two current waveforms, one in the linear region and one in the saturated region. The current spike occurs because the flux linkage continues to increase; but in the 78 saturated region, the inductance is very low, necessitating a rapid increase in current to match the volt-seconds applied to the winding. This current increases until the IR drop matches the applied voltage. The actual current waveform would be smoother than depicted in the model, owing to the model's piecewise linear assumption. If the applied voltage across the winding minus the voltage drop due to the winding resistatnce is integrated over time, the flux linkage is obtained. Plotting the flux linkage and current on separate axes, the cyclic flux linkage-current profile is obtained as shown in Figure 3-11 part (b), where the saturated and unsaturated profiles correspond to the current waveforms plotted in part (a) of the figure. The area within these loops is the net work per cycle. The flux linkage-current profile (net cycle work) is determined by the turn on, 0"", and turn off, 9Of, angles. Determining the optimum excitation is dependent upon the balance between generating larger forces with larger current and dealing with the PR losses associated with those higher currents. 3.2.3 Loss Mechanisms The efficiency of the transducer can be estimated by modeling the various loss mechanisms. The three primary loss mechanisms are the winding, eddy current, and hysteresis losses. The inverter losses are neglected for this analysis. Typically, the eddy current and hysteresis losses are significantly smaller than the winding loss and are lumped together as core losses. Early in the design of this generator, neglecting the effects of eddy currents resulted in poor design decisions. The original design used a ferritic stainless steel stator and piston in an axisymmetric geometry, which would be simple and relatively inexpensive to fabricate. Although this design appeared attractive because of its simplicity, it was determined impractical due to the high eddy current losses, even though the electrical resistivity of ferritic stainless steel is approximately 50% higher than typical silicon iron or iron cobalt laminations. The skin depth for the steel is given by 6= 2psleel W7 79 (3.28) where Pateel is the resistivity of the steel. This led to a calculated skin depth of approximately 0.5 mm, which indicated that the thickness dimensions of the steel should be smaller than 1 mm to avoid eddy current losses. This was not feasible for the axisymmetric design, and therefore, the generator was redesigned using laminated electric steel components. The generator built used 29 Gauge M19 electric steel to minimize the eddy current losses. Higher flux carrying materials such as iron cobalt (Hyperco) laminations could be used to increase power density but at a significantly higher price. The fabrication of the generator is discussed further in Section 3.5. The core loss is difficult to determine analytically. Typically, the core loss can be estimated by the power loss per kilogram of laminated steel based on the frequency of operation and maximum flux density. However, applying the core loss from the published B-H loop data to the generator in this research overestimates these losses because published core loss data assumes that the magnetic flux in the core switches direction. In this VRG design, the flux does not change direction since the magnetic force direction between the stator and piston components does not depend on the flux direction. The entire hysteresis loop experienced by the VRG core extends only in the positive magnetic field intensity (H) range. Consequently, the total core loss (proportional to the area of the hysteresis loop) is smaller than what would be exhibited by a machine that drives the core over the entire range of magnetic field intensities. The relative sizes of the hysteresis loops for complete magnetic flux reversal and the minor hysteresis loop associated with the VRG in this research are shown in Figure 3-12, where the shaded region represents the area within the minor hysteresis loop. An estimate of the core loss can be obtained by applying Pe =] W2 t2 fvoi p eel dB I - | dV d (3.29) to the model of the VRG [30], where tiam is the thickness of the laminations. The approximate flux density is calculated at each time step of the VRG simulation with the equation B - NA 80 (3.30) aI Figure 3-12: This figure depicts qualitatively the minor hysteresis loop traveled for the unipolar excitation of the core material. This leads to less hysteresis losses than in flux reversing cores. Figure taken from [30]. where A is the steel's cross-sectional area. Since the cross-sectional area of the piston steel is half that of the stator, the flux density was calculated separately for the stator and piston and then summed to determine the total core loss. These equations underestimate the core loss for saturated steel but was included in the model to get an estimate for the core loss and more accurately optimize the generator system. The gauge of the motor wire and number of turns for determining the wire resistance is directly tied to the geometry of the generator shown in Figure 3-3. These variables are only relevant when determining what voltages and currents should be used for driving the VRG. The choice of wire gauge and turns were determined by selecting drive circuitry components that are within the range used in the automotive industry to keep the electrical component costs low. This will be discussed further in Section 3.3. The resistance of the wire is R = pwLw Aw (3.31) where pw is the wire resistivity, Lw is the length of the wire, and Aw is the area of a 81 single wire. The approximate length of the wire was calculated by LW = 8lwhw(li + 1w)( A 2 (3.32) This gives the length of wire, based on the volume available, for the coil compared to the volume of the coil corrected by a packing factor, (. The ideal packing factor for windings stacked directly on top of each other is pi/4. This is the value that was used for the analysis. The number of turns is then N = (Wh A (3.33) W Solving these equations together gives an approximate resistance for the coil; R 8(l + E')N 2 (3.34) 2 The winding losses are then calculated for the two generator phases such that 2 Pw = E RI2 (3.35) n=1 where In refers to the current of phase n of the two phase generator. The winding losses are the most significant loss contribution particularly in high power applications. 3.2.4 Flux Tube Analysis and Fringing For most generators, particularly the rotary type, the air gaps are very small throughout the duration of the cycle. This leads to nearly negligible fringing effects. Due to the large air gaps of this engine, it was determined that fringing would start to have a significant effect particularly at larger air gap lengths. The addition of fringing into the model effectively raises the minimum inductance. This increase in minimum inductance can decrease the work output per cycle depending on the excitation. For this reason, fringing was included in the model so that the power output capability 82 Piston Steel Outer Air Gap RFringe Inner Air Gap Reap RFringe RGap NI coil Stator Figure 3-13: This figure depicts the magnetic circuit used for the more detailed analysis including fringing and leakage flux paths. The parallel reluctances RFringe and RLeakage have a significant effect on the magnetic circuit particularly at larger air gaps. would not be overestimated. The addition of fringing and leakage fields create par- allel reluctances in the magnetic circuit analogy. The magnetic circuit used for the flux tube analysis is shown in Figure 3-13. The flux tube analysis done for this generator follows the work of Herbert Roters [18]. The method for calculating the fringing and leakage fields is based on what Roters calls estimating the"permeances of probable flux paths." The permeance is simply the inverse of reluctance. The solutions are geometric in nature, and having established the generator geometry, the probable flux paths could be calculated using the same geometric variables as given in Figure 3-3. The air gaps were separated into an inner air gap and an outer air gap, and each air gap was broken into simple-shaped volumes that contain all of the potential flux paths. These simple-shaped volumes have permeances that are easily approximated. The volumes used for the flux tube analysis include partial cylinders, partial annuli, spherical segments, and quadrants of spherical shells. An example of two of these flux path volumes is provided in Figure 3-14, where the half cylinder represents the flux from edge AB to edge GF, and the half annulus represents the flux from surface B to surface F. In total, 15 different 83 H E Figure 3-14: This figure shows two of the simple-shaped volumes used for the flux tube analysis. The half cylinder represents the flux from edge AB to edge GF, and the half annulus represents the flux from surface B to surface F. permeances were calculated for the possible flux paths between the air gaps, and the permeance of the leakage path was also determined. The leakage flux is the flux path from surface "H" to surface "C" in the figure which completely bypasses the piston. Appendix A includes the equations used for determining the different permeances, which change as the air gap length changes. The permeances for the outer air gap and inner air gap were summed in parallel to ultimately obtain values for the inner and outer air gap fringe reluctances labeled RFringe in Figure 3-13. The reluctance, RGap, is the reluctance of the flux path perpendicular to the stator and piston air gap surfaces labeled "A" and "G" in Figure 3-14 for the outer air gap and "I" and "G" for the inner air gap. The addition of fringing and leakage fields tends to decrease the current at which saturation occurs in the laminated steel. This can be seen clearly when looking at a plot of the magnetic forces as a function of displacement at constant current. This is shown in Figure 3-15 part (a). The effect of the fringing and leakage fluxes can be seen by comparing Figure 3-10 part (a) with Figure 3-15. 84 0 Force vs Displacement at Constant Currents v-s0.045 Fringing Effects on Flux Linkage-Current Plots 0.04-500. 0.035 0.03 -100 0.025 S=34 1=5A -l=7A -=9A 3 0.02 -1=11A LL 0.015- -- S-150 0 -200 I=1 3A - 1=15A -250 0.0050 1 w/o Fringe Effects W/ Fringe Effects 5 Air Gap Length (m) x 10- 10 15 Current (A) (a) (b) Figure 3-15: This figure shows (a) the force at constant current for the range of air gap lengths. This figure should be compared with Figure 3-10 to see the effects of fringing and leakage fluxes. (b) Shows the effect of fringing and leakage fluxes on the flux linkagecurrent plots. The inductance is increased as the total reluctance of the flux linkage path has decreased. This effect is particularly pronounced at larger air gaps and limits the work output per cycle. Excluding the fringing and leakage would overestimate the power output capability of the generator because the magnetic forces are significantly smaller at larger air gaps than would be predicted otherwise. Figure 3-15 part (b) shows the effect of fringing and leakage fields on raising the minimum inductance. Therefore, the flux linkage-current profile will enclose a smaller area and thus have a lower work output per cycle. The results of this analysis led to the conclusion that to produce an accurate model of the VRG, the effects of fringing and leakage fluxes should be included due to the potentially large air gap lengths inherent to the design. 3.3 Power Electronics There are multiple configurations for excitation of the winding. There are trade- offs for each design as described in references [6] [28] and [30]. For the design of this VRG, only two phases exist, and thus, the number of MOSFETs was not an important 85 Gate Driver, Gate Driver,J Wnding 2 Winding 1 JLGate Drlverj A Gate Driej Figure 3-16: This figure shows a schematic of the drive circuitry used for the VRG. The two windings are driven 180 degrees apart during the piston cycle. consideration. Additionally, the simplification of having only one power supply and only one winding led to the implementation of the drive circuitry as shown in Figure 3-16, where the gate driver for the MOSFETs provides the on/off signal and ensures fast switching to reduce circuit losses. Only one phase of the VRG was used for experimentation as will be described in Section 4.2. The necessary power electronics were placed on a printed circuit board (PCB). The PCB was specifically designed to limit the physical loops between the gate driver and high-side/low-side MOSFET transistors. This is to prevent any issues with ringing in the circuitry while switching the power MOSFETs on and off. The schematic used for the PCB is shown in Figure 3-17. A number of considerations went into the design of the drive circuitry. A bootstrap circuit was used to maintain the charge on the high-side MOSFET. Additionally, a fuse, labeled F1 in the figure, was included to ensure if the VRG were driven too Table 3.2: Drive Circuitry PCB Components Circuit Components Gate Driver (MDr) Resistors (RI, R2, R3) MOSFETs (MOS1, MOS2) Diode (D1) Diodes (D2, D3) Capacitors (C1, C2) 86 FAN7360 1OQ NDB5060L SK31OA-LTP SBR15U50SP5-13 3pF, 1pF Fl30- ~ 1 41t ....... L s.. RE Figure 3-17: This figure shows a schematic of the drive circuitry used for testing the VRG system. The winding was soldered across connections labeled W2 and W3. A LabVIEW pulse signal was applied to the connector labeled Coni. More detail on the experimental setup is described in Section 4.2. heavily into saturation and large currents were experienced, the circuit components and measurement equipment would not be damaged. The coil ends of the VRG were soldered to points W2 and W3 as labeled on the schematic, and traces that - experienced the same currents as the winding were made between 5 - 6.5 mm (0.2 0.25 in) wide. Table 3.2 lists the components used for the drive circuitry. These components were specified for the original operation of 250 Hertz with a 50 volt driver and expected peak currents up to 15 amps. The number of turns in the coil and the size of the wire were targeted to keep the voltage and amperage ratings of the MOSFET components within the automotive industry's standard. Without considering the limitations of electrical components, the design optimization may provide unrealistic current profiles or applied voltages. This is especially true for higher frequency operation because of the large voltages necessary to quickly drive current into the windings. There are potentially three stages to motor excitation: charging, freewheeling, and 87 Charging Freewheeling Generating Figure 3-18: The three excitation stages are the charging of the winding, freewheeling, and generating. The freewheeling stage was not included in the analysis of this VRG. The charging stage drives a small current into the winding and the generating stage drives a larger current back into the power supply delivering a net electrical power output. generating. These three stages are depicted in Figure 3-18, where the charging stage occurs when both MOSFETs are operated as closed switches, the freewheeling stage occurs when the high side MOSFET is an open switch but the low side MOSFET is a closed switch, and the generating stage occurs when both MOSFETs are open switches. The charging stage is used to drive current into the winding and provides input to what is essentially a parametric amplifier, driving more current out of the winding during the generating phase. The use of the freewheeling stage was not considered for this work. Other potential excitation schemes and methods for determining optimum excitation are described in [6] and [28]. 3.4 Design Optimization Geometric optimization was completed for the VRG by calculating the power output and efficiency of the generator and performing a grid search method for the geometric variables provided in Figure 3-3. A number of variables were fixed, including the outside diameter of the generator (.0762 m) and the inner and outer air gap areas were kept equal. The dimension "z," as labeled in Figure 3-3, was held constant at 0.25 mm to keep individual laminations from having thin long sections, which would greatly increase fabrication difficulty and cost. The h, dimension was also left fixed for structural purposes and was significantly thicker than the piston steel 88 so that saturation limitations would not be an issue in this segment of the stator core material. This left only 3 geometric optimization variables: h", 1j, and hp. The design choices were based on the transduction efficiency, normalized power output, and inverter rating. The inverter rating being important to preclude unrealistic or costly drive circuit components as discussed in the previous section. The efficiency of transduction is defined as IPO 1P01+ jPWI + IN (3.36) where PO is the power output of the generator, Pw is the power loss due to winding resistance, and P. is the core loss. For each set of VRG power performance calculations as a function of coil height, piston lamination height and air gap dimensions, there exists a peak power point. Since the power and efficiency trends are of the most interest here, the results are graphically presented using normalized power output, PO* defined as P* = where IPol I(3.37) max[|Po|] is the power output and max[Pol] is the power corresponding to the maximum power point of the calculated curve. The variable h, represents the maximum height of the coil winding. As the height of the coil increases, the efficiency increases because the same number of turns can be made with larger and larger diameter wires, thus reducing the PR losses. However, because the leakage flux was included in the model, there exists a maximum power output because as h, increases the reluctance of the leakage path (described in Section 3.2.4) decreases. This reduces the useful flux traveling through the air gap to the piston, which provides the magnetic attraction forces between the piston and stator. The normalized power and efficiency curves as a function of the winding height are shown in Figure 3-19, which were produced by fixing the excitation angles and all other variables except the winding height. The chosen design dimension lies between the maximum efficiency point and the maximum power point. The winding height becomes more relevant considering the power density of the device because of the 89 Design Optimization Curves for Coil Height Dimension 0.95 0.9 0.9 0.85 0-- -M 0.8 Efficiency -Normalized 0 Z 0.7 0.75 :Chosen Desin - 8. Power 10 16 14 12 18 20 22 24 Length Dimension for Coil Height, hw (mm) Figure 3-19: This figure shows the chosen design point for the coil height based on the efficiency and normalized power output of the generator. The power output is normalized such that the maximum is set to one. significant weight of copper windings. The variable 1i is a more complicated variable in terms of optimization. All of the air gap dimensions are driven by this length dimension. To get an accurate prediction of the optimum air gap design, the volume of magnetic material in the piston was fixed and the length dimension, 1j, was varied. The volume of magnetic material in the piston had to be fixed in order to decouple the piston saturation effects with the air gap geometry. A value for this dimension was again chosen between the power and efficiency maximums. The results of this design study and the selected design dimension are shown in Figure 3-20. The larger the mass/volume of the piston steel, the larger the cross-sectional area available for the magnetic flux, thus increasing the current at which the steel saturates. Increasing the piston dimensions increases the overall piston mass, but can also increase the power output capability of the system. The same design study for the dimension 1i was repeated, but the piston mass was allowed to increase. The results are exactly as expected and shown in Figure 3-21 part (a). The normalized power output of the generator increases until the point again where leakage flux begins to dominate. However, shown in part (b) of the figure, the necessary voltage to drive 90 Design Optimization for Stator Air Gap Dimensions 0.9- - 0.8.- - Efficiency Normalized Power 0.70 (- 0.60 Z :Chosen Desin 0.5- 0.4 1'0 11 12 Driving Length Dimension for Stator Air Gap Geometry, Ii (mm) Figure 3-20: This figure shows the chosen design point for 1i length dimension, which was set as the driving dimension for the air gap geometry. The piston volume was held constant, and the current waveform was kept constant by driving at the same turn on and turn off angles and peaking at the saturation current. the VRG also increases because the overall inductance of the generator increases with larger piston steel volume, which complicates the drive circuitry and increases the required inverter rating. The maximum drive voltage selected was 50 volts and peak currents of 15 amps. The importance of the inverter rating was overestimated in this research, and the drive voltages should be reconsidered for the power electronics developed for the complete portable power system. A value of 3.8 mm for the piston height, hp, was chosen to reach the desired 50 watt output with the stated voltage and current limits, but kept as small as possible to keep the mass of the piston low, which is desirable for higher frequency operation as discussed in Section 2.2. The power output of the generator could be increased by simply increasing the dimensions of the piston steel. Apart from the geometric optimization, the turn on and turn off angles significantly affect the efficiency and power output of the generator. The geometric optimization was done around a single turn on and turn off angle based on an approximate peak in power and efficiency regardless of the geometric dimensions. For experimental validation of the model, the VRG was tested at different turn on and turn off angles 91 Excitation Voltage for Stator Air Gap Design Optimization for Stator Air Gap Dimensions 0.95- Optimization 90 -- - 4) 0.9 80 0 4.8> 0.8- C _ oi Dimension 0 - .Efficiency ccNormalized Power Z 70- Chosen Design 60 0.7 -0.65- 'Chosen Design 50 Dimension 6 8 10 12 14 16 Driving Length Dimension for Stator Air Gap Geometry, li (mm) (a) 6 8 10 12 14 16 Driving Length Dimension for Stator Air Gap Geometry, li (mm) (b) Figure 3-21: Shown in (a) is the effect of increasing the air gap dimensions and subsequently decreasing the available volume for the coil. The optimum is shifted when the volume of the laminated steel piston components is allowed to increase. (b) Shows the increase in applied voltage necessary to drive the higher inductance generator to the same steel saturation point. to determine optimum power and efficiency points. This data as well as the model determination of the optimum turn on and turn off angles is presented in Section 4.3. 3.5 3.5.1 VRG Fabrication Laminated Components The mechanical design incorporated the model optimizations and was done with an emphasis on the final product, which will ultimately be the portable thermoacoustic engine. As previously discussed in Section 3.1 and Section 3.2.3, the core material had to be laminated steel to suppress significant eddy currents, which would otherwise dominate the system. This led to the design of the laminated steel wedges. All laminated components were manufactured by Polaris Laser Laminations. The number of wedges was chosen so that the 90 degree cuts made by the laser cutter would align during assembly of the wedges to make a single component. This worked to an extent, 92 (a) (b) Figure 3-22: Shown in (a) are the materials used for assembling the stator, and in (b) is a closer view of the stator pieces as they were set in the delrin jig. but the tolerance on the 45 degree wedge angle left a small gap where the wedge faces met in the center of the inner air gap. Engineering drawings for all components used in the VRG are provided in Appendix D. The dimensional accuracy held to the drawings for the laminated steel components was 0.025 mm for critical dimensions. The thickness, as defined by the dimension "w" in Figure 3-3, was within 0.15 mm". The poor dimensional toler- ancing of the thickness dimension made fabrication of the VRG more difficult and should be more precisely accounted for in future designs. The four stator pieces were then assembled using Stycast 2850 epoxy and a right angle delrin jig fabricated for this purpose as shown in Figure 3-22. The epoxy was placed on the wedge faces and then pressed together. A urethane release agent, Camie 980, was used on the jig to ensure any epoxy that left the wedge faces would not adhere to the jig. Delrin was used for the jig material because it has a natural resistance to the epoxy adhesion. To ensure parallelism and flatness of the air gap surfaces, the surfaces were held against a smooth glass surface. The epoxy was cured in an oven at 200'F for 2 hours. A flat plate of aluminum was used as a weight to ensure the air gap surfaces were held against the glass. The rubber gasket material was placed between the aluminum and the stator pieces to ensure even distribution 93 Aluminum Plate Gasket Material Stator PiecesC-lm -lm D elr in Glass r ac Surface Epoxy Sample Figure 3-23: This figure shows the stator assembly fixture system as it was baked to cure the epoxy. The C-clamp was necessary for the fabrication of this stator to hold an accurate stator width dimension. of force as shown in Figure 3-23. Type 316 stainless steel brackets were made to hold the stator as shown in Figure 3-24. These brackets, in the complete system, protrude through the piston ports, allowing the stator to be held internally to the oscillating piston as discussed in Section 3.1. The stainless steel brackets were annealed to remove residual magnetism left from the machining process. In the pre-annealed state, a small magnetic attraction could be observed when a magnet was held against the brackets. The brackets were brought above 10000C using a propane torch. This method caused slight oxidation of the stainless steel, but was a very quick method for getting rid of residual properties when it was unimportant whether it oxidized or not. The post-annealed brackets revealed no detectable magnetic attraction. This was done to prevent any magnetic flux from permeating the stainless steel bracket, which would result in eddy current dissipation due to its relatively large thickness dimensions compared to the laminated steel sheets. Unfortunately, because of the poorly toleranced overall thickness dimension of the laminated stator components, brass shims were necessary to fit between the set screw and outermost steel laminations. This had the added benefit, however, of allowing the set screws to be set with a large amount of force without concern of damaging 94 Outer Air Gap Surface Stator Bracket V Shims Figure 3-24: This figure shows the stator assembly with all components except the windings. Glass Surface Gasket Material Stator Assembly Piston Laminations Piston Laminate Fixture -"Piston Cap" (a) (b) Figure 3-25: Shown in (a) is the method for ensuring parallelism between the stator and piston laminations. Not shown in this figure is the gasket material which was placed between the piston laminations and the piston cap. (b) Shows the piston laminations after assembly. the laminations. Thread-locking set screws were used to prevent vibrations from loosening the brackets over time. One design change that should be implemented is the use of one additional set screw on each side of the bracket in order to prevent any rotational motion of the brackets. For this design, the steel laminations overhung the bracket to prevent the bracket from sliding. This significantly strengthened the brackets from sliding axially, but shims were again necessary above and below the bracket because of the 0.025 mm tolerance. It was decided to not attempt a press- fit for the brackets, which may have damaged the laminations or affected the more important stator length dimension. 95 The piston laminations were set into an aluminum fixture. In the full design, this aluminum, referred to as the piston cap, would provide the continuous surface necessary for the thermoacoustic gas spring pressures to act on as discussed in Section 3.1. The aluminum fixture shown in Figure 3-25 part (a) is over-sized relative to the full piston design and was used for experimental testing of one half of the VRG system. To ensure optimum parallelism with respect to matching stator and piston components, strips of 1.5 mm rubber gasket material were placed between the laminated piston components and the aluminum holding piece. The stator laminations were then aligned to the piston laminations and weighted so that the compliant elastomer allowed the piston to exactly conform to the stator air gap faces as shown in part (b) of Figure 3-25 while the epoxy set and fixed the laminations in place. Rubber gasket material was again used for pressure application above the stator assembly to ensure even distribution of force on the piston laminations. Post fabrication analysis using feeler gauges revealed that the minimum effective air gap length was 0.1 mm" when the stator sat directly on the piston laminations. 3.5.2 Coil The coil was wound using a mandrel fabricated out of nylon. Nylon was used to prevent the adhesion of the epoxy, and was used in combination with a urethane mold release. It was estimated that 160 turns could be made out of 16 gauge motor wire, but ultimately 148 turns were all that could be fit to preserve a modest clearance between the coil and laminated steel components. With more experience and precision winding methods, the 160 turns is definitely attainable considering the gap that was left between the coil and laminations. The coil was turned slowly on a lathe as shown in part (a) of Figure 3-26, and a foot operated switch was used to manually turn the lathe chuck on/off. A cycle counter was fixed to the end of the lathe to keep track of the number of revolutions. The coil was wound while Stycast 2850 epoxy was simultaneously applied to each layer of the coil. This particular epoxy was chosen because of its excellent thermal properties. After completing the winding process, a nylon sleeve was placed over the coil to 96 (b) (a) Figure 3-26: Shown in (a) is the lathe and a completed coil with the nylon cover over the winding. (b) Shows the winding after the epoxy has cured and the cover has been removed. The coil shown was a first attempt. In subsequent attempts, additional epoxy was used leading to a significantly more continuitous epoxy coating. Figure 3-27: This figure shows the stator assembly with one of the two phase coils. This assembly was used for experimental validation of the VRG. ensure that the coil was within the allowable dimensions set by the laminated stator assembly. The lathe was then run continuously for several hours to ensure the epoxy was uniformly distributed throughout the coil and did not shift due to gravity. After 97 16 hours of curing, the nylon cover was removed and the mandrel was separated into its two pieces. The coil and mandrel are also shown in Figure 3-26 part (b). The coil was then placed on the stator and epoxied in place with 2850 epoxy. Only one side of the VRG was necessary for the experimental validation. In the full generator, a second coil would be placed on the opposing side of the stator. The second coil would be wound in a manner such that magnetic flux is driven in the same direction as the first coil. Otherwise, significantly larger core losses will be experienced as the full hysteresis loop will be traced in sections of the stator. The full assembly of the stator used for experimental validation is shown in Figure 3-27. The resistance of the coil was measured using a Hewlett Packard 34401A multimeter and a resistance of R = 0.253 Q was found. This value was subsequently verified in future experimental tests by measuring the applied voltage and current in steady state. 3.6 Chapter Summary A design model of the VRG that accounts for saturation effects of the laminated steel, the leakage and fringing fluxes, and the losses due to both resistance in the coil and the core losses was generated. This model allowed the prediction of current waveform and drive voltages that could be used to specify and fabricate the required power electronics. The losses associated with the electronic components including the diodes and MOSFET transistors was not included in the model. The model was then used to optimize the specific geometry of the VRG system using laminated steel components. Based on the optimization results, structural, and power output requirements, the stator and piston were designed and fabricated. Chapter 4 provides the experimental validation of the model and the overall VRG system. 98 Chapter 4 Experimental Design Verification The models developed in the previous chapter were experimentally validated in two ways: a static analysis for generation of flux linkage-current plots, and a dynamic test to verify the power output projections of the model. Section 4.1 describes the experiment setup and data collected for the generation of the flux linkage-current plots. Section 4.2 describes the experiment design used for the model verification, and Section 4.3 presents the results of the dynamic experiment. 4.1 Saturation Characterization The static experiment was conducted to produce the flux linkage-current characteristics of the generator at fixed air gap lengths. The air gap is the free-space between the piston face and stator pole as shown in Figure 3-5 labeled as the inner air gap and outer air gap. Displacement of the piston causes an an equivalent increase or decrease in the air gap. The flux linkage-current plots are necessary for understanding the magnetic forces and work output capability of the generator as discussed in Section 3.2. To generate these plots, the generator stator and piston were aligned, and plastic shims were used to accurately gauge the air gap thickness as shown in Figure 4-1. A voltage was then applied across the windings and the winding current and voltage were simultaneously measured and used to determine the flux linkage-current characteristics at each air gap length. 99 Stator Winding Plastic Shims Piston F _____________________ Laminations Figure 4-1: This figure shows the setup used for generating the flux linkage-current plots for different air gap lengths set by the plastic shims. To ensure accuracy of the measurements sources of potential error were identified and mitigated if possible. Plastic shims (orange/yellow in Figure 4-1) were necessary to prevent the generation of eddy currents in the shim material, which would cause the maximum inductance to appear larger than the true maximum inductance. Another source of error for the experiment may have come from elastic deformation of the plastic shims due to the magnetic forces The elastic deformation would tend to decrease the thickness of shims to less than their stated thickness. The error due to elastic deformation was calculated to be less than 1% by assuming a 275 Newton compression distributed across the air gap surfaces. This error was considered negligible and was subsequently neglected. If the shim materials were not initially flat, the air gap would be larger in some sections than the stated shim thickness until the magnetic attraction forces between the piston and stator poles flattened each gap to exactly the shim thickness. To minimize the effect of "bowed" shims, approximately 10 pounds of pre-load was applied to compress the gap uniformly to the shim thickness. With no shim in place, the stator poles rest directly on the face of the piston laminated pieces. However, due to small deviations in the laminations themselves or limitations in the manufacturing process, the plane formed by the stator poles is not directly parallel to the plane formed by the faces of the piston laminations. For 100 Gate Driver 2-1 Winding Resistance Supply___V V. Phane Winding Figure 4-2: This figure depicts the circuit used for driving the currents through the winding to measure the flux linkage-current characteristics at various air gap lengths. The diode was necessary to allow the current to freewheel after the MOSFET "switch" was opened to prevent charge buildup and extreme voltages across the MOSFET component. this reason, the piston and stator were placed in direct contact, and the resulting gap between each stator pole and piston face was measured using feeler gauges. The measured gap was found to be up to 0.1 mm. The tolerance of the laminations themselves were +0.025 mm, which indicates the fabrication process described in Section 3.5 potentially added an additional 0.075 mm to the gap between the stator and piston. This gap dimension also represents the smallest possible air gap used in the calculation of the air gap reluctance. This corresponds to the maximum inductance and is referred to as the zero air gap in this document. This is different than the term minimum air gap, Gmin, which is used in Sections 4.2 and 4.3 to refer to the minimum gap achieved during the cyclic motion of the piston as it moves closer and further from the stator. The circuit shown in Figure 4-2 was used to drive the current through the winding at fixed gap displacements. A LabVIEW program was used to send a single pulse to the gate driver, which subsequently drove the MOSFET to act as a closed switch. Current was then driven through the winding by the power supply until the LabVIEW pulse ended and the MOSFET switch was "opened". Due to the inductance of the winding, the current in the coil then freewheeled around the path constituting the diode and winding until the current was dissipated in the winding. The same gate driver, MOSFET, and diode were used for both the static exper101 Current Profile for Flux Linkage-Current Plots Voltage Profile for Flux Linkage-Current Plots 6.5 12- 6- 10 5.5 Interal Impedance Voltage Drop 8 5-- 4.5. 4 4- 2 3.5 0 0.002 0.004 0.006 0.008 Time (s) 0.01 0.012 0.014 0 (a) 0.002 0.004 0.006 0.008 Time (s) 0.01 0.012 0.014 (b) Figure 4-3: This figure depicts (a) the applied voltage to the phase winding as a function of time. The integral of this voltage minus the IR component provided the flux linkage. (b) The current profile for the static test at zero shim gap is shown. The current changes slowly until the steel begins to saturate at about 6 ms. iment and the dynamic drive circuitry presented in Section 3.3. An HP 6428B DC power supply was used to drive up to 14 amps of current through the winding at 6.38 Volts. The voltage was chosen so that the maximum current measured would be approximately 20% greater than the expected peak current for the dynamic operation of the generator. Additionally, the voltage drop due to the resistance of the winding was less than the inductor emf driving voltage for all currents. Therefore, the experiment was terminated when the voltage drop across the winding was approximately equal to the IR drop, where I is the current and R (0.25 Q) is the winding resistance. A Tektronix MSO2014B oscilloscope with a passive voltage probe and a TCP0020 current probe were used to simultaneously measure and record the voltage and current. The current data was put through a low-pass (fcutoff = 1 kHz on 300 kHz sampled data), zero-phase shift filter to smooth out the high frequency components due to the ADC resolution of the oscilloscope. The voltage applied minus the IR voltage component was integrated to determine the flux linkage. Figure 4-3 shows the voltage and current profiles typical for the experiment. It can be seen that the 102 Flux Linkage-Current Plots 0.07 0.06r - 0" 0.001" - 0.002" - 0.004" 0.0075" 0.01" - 0.05 - 0.015" 0)0.04 0.02" 0.025" 0.03" - 0.03 - 0.04" - 0.05" - 0.06" -- 0.02 - - 0.08" 0.1" 0.12" 0.15" 0.18" 0.22" 0.01 U O 0 2 4 6 8 en 2p4 6 Current (A Figure 4-4: This figure shows the flux linkage-curr ent plots at a range of air gap displacements measured experimentally. voltage measurement is significantly more noisy than the current measurement, but this does not significantly alter the flux linkage measurement due to the smoothing effect of the integration. Also shown in part (a) of the figure is a slight voltage drop (6.4 -> 4.8 V), which can be attributed to the internal impedance of the power supply. Plotting the integrated voltage data with the filtered current data at every point led to the flux linkage-current profile for a single air gap length. Data was collected for 37 air gaps lengths, 19 of which are shown plotted in Figure 4-4. As predicted, the steel saturates at approximately a constant flux linkage instead of a constant current, which can be seen in the figure noticing that each line becomes nonlinear at approximately the same flux linkage value (0.038 Wb). A comparison of the actual flux linkage-current plots to that of the model revealed that at larger air gaps the model deviated substantially from the actual data. The model predicted a minimum inductance (LMin = 1.75 mH with air gap G = 5.6 mm) that is significantly less than the experimentally measured minimum inductance 103 Flux Linkage-Current Plots Predicted vs Measured F 0.08- 0.08--- ---Predicted rd.ed 0.07- ---Measured -- -Measured 0.04 0.07- 0.07 -** 0.07 -d - 0.05 0.05 c-0.02 0.01 Flux Linkage-Current Plots Predicted vs Measured 0.0 * 0.0 0.02 -- J.0.01 0 10 15 Current (A) (a) 0 5 10 15 Current (A) (b) Figure 4-5: This figure depicts (a) the original model compared to the experimentally determined flux linkage-current plots. The model significantly underestimates the magnitude of the fringing fluxes. The flux linkage plots are for gaps ranging between 0 - 3.81 mm. (b) This plot shows the model with the permeances of the flux path increased by 40%, which far more accurately models the true flux linkage characteristics. (LMi, = 2.6 mH with air gap G = 5.6 mm). This suggests that the effect of the fringing fluxes is larger than what was predicted by the flux tube analysis. Increasing the effective permeance of each flux path by 40% allows the model better match the measured flux linkage characteristics of the generator. These results are shown in Figure 4-5, where part (a) of the figure shows the experimental data and the model predictions without the fringing correction factor and part (b) shows the experimental data and the model predictions with the fringing correction factor. The accuracy of the model is significantly improved in the linear region as evidenced by the closeness of the predicted and measured lines in the figure, but there are clear limitations of the piecewise-linear approximation in the saturated region. The rounded saturation region is poorly matched by a single saturated incremental inductance, which leads to inaccuracies when predicting power output, efficiency, magnetic forces, and current waveforms using the piecewise-linear model. It is now possible to create an improved model for predicting the generator charac- 104 teristics using the measured flux linkage-current data. The data can be fit and a new model generated based on the actual flux linkage-current characteristics of the generator. This approach is described in detail in [29]. From this model, the power output, forces, and efficiency can be predicted with greater accuracy demonstrated in [29]. The piecewise linear model is useful for making the correct design decisions before the flux linkage-current data is available, but now that the data is available, it is left as future work to generate a new model based on the measured flux linkage-current data. To verify the accuracy of the data collected, the approximate B-H curves for the steel used in the magnetic core were generated based on the flux linkage-current measurements. This was done by using the zero air gap static test, where the piston steel was placed in direct contact with the stator poles. In this case, the flux loop is nearly continuous but with a small additional air gap reluctance due to the manufacturing limitations of the magnetic core. The B-H curve produced by this method is only an approximation because the cross-sectional area of the flux loop is not constant in this design of the generator core and the small unavoidable air gap still has a significant effect on the flux loop compared to that of a continuous toroid. The flux density, B, and magnetic-field intensity, H, are given by the equations A B (4.1) NA H = (4.2) H1 where A is the cross-sectional area of the steel, lc is the mean core length, N is the number of turns in the coil, A is the flux density, and I is the current in each wire of the coil. The B-H curves calculated by this method and published data for M19 29-gage steel are provided in Figure 4-6. The green plot represents the calculated values for the B-H curve assuming the magnetic loop has everywhere the cross-sectional area of the stator (A = 2.71 i 2 ). The red plot represents the calculated values for the B-H curve assuming the loop cross-sectional area is that of the piston (A = 1.63 2 ), and the blue plot is the published B-H curve data. As expected, the actual B-H saturation for 105 = -R) Comparison of 0 Air Gap B-H Data to Published Data 3 2.5- 2 1.5- 0 - 0.5- 0 4000 8000 12000 16000 H (A/m) Figure 4-6: A plot of the calculated B-H curves based on the zero air gap test of the flux linkage-current plot for both the maximum and minimum cross-sectional areas of the laminated generator components. The data cannot be fit directly to the published data because the cross-sectional area is not constant, and therefore, the published data lies at some mean cross-sectional area. Additionally, the approximate minimum air gap can also be calculated based on the difference in inductance between the published and acquired data. This minimum air gap was calculated to be 0.066 mm, which is reasonable considering the physical measurements done for this zero air gap discussed earlier in this section. Published curve from [9]. published data lies somewhere between the maximum and minimum cross-sectional areas of the generator data because the saturation is ultimately affected by both steel sections. It is also possible to calculate an approximate minimum effective air gap when the stator and piston are pressed directly together. The inductance in the linear region of the B-H curve is relatively unaffected by the cross-sectional areas. Therefore, the inductance of the published data and the experimental data were calculated, and using Equation 3.6, the air gap reluctance was also calculated. The approximate air gap length is then given by (4.3) 32 N L where Ag is the area of the inner and outer air gaps and R is the reluctance of the 106 steel as calculated from the inductance of the published data. The approximate minimum air gap calculated from this analysis is 0.066 mm. This value fits well with the measured 0.1 mm air gap using the feeler gauges, and is reasonable considering the laminations themselves were manufactured with dimensional tolerances of +0.025 mm. From this analysis, it was concluded that the static experiments were accurate and the data could ideally be used for more accurately modeling the dynamic performance predictions for the generator. 4.2 Experiment Design To experimentally validate the electroacoustic transducer, the added complexity and control of the thermoacoustic engine was removed. The initial proposed design for testing the VRG at the 250 Hertz operating point was to set up a resonant system using two gas springs and the center piston, driven to resonance by a shaker table. Mechanical springs are not an acceptable substitute for gas springs because the spring constants required for 250 Hertz vibrations and 5 mm peak-to-peak oscillation amplitudes of a 0.2 kg piston are on the order of 250, 000 N/m. The size of the spring necessary for this spring constant creates a situation where the spring mass is as significant as the piston mass for the oscillation. The resonant frequencies are very difficult to obtain in this situation. For this reason, it was determined that the gas spring system was necessary for testing the electroacoustic transducer. The resonant system was necessary to test the VRG system at the proposed 250 Hertz operating frequency, because the shaker table cannot directly drive the generator at the required amplitudes at that frequency. However, without a gas bearing system, another type of bearing and clearance sealing method would be required. This added complexity was not feasible for the time-scale of this project and was not necessary for initial validation of the VRG system and VRG model. For this reason, a simpler and less costly experiment was developed to test the VRG by directly driving the VRG system with the shaker table. With this method only one phase of the two-phase generator was tested and at a significantly lower frequency than for 107 the original design operating point. The lower frequency experiments were ultimately used to test the model, the power output of the generator and the efficiency. The final power output and efficiency could be reasonably determined by scaling the experimental data to higher frequency using the known loss mechanisms as discussed in Section 4.3.1. The dynamic tests were done such that the stator was fixed above the shaker table, and the piston laminations were driven with a sinusoidal motion by the shaker to change the inductance of the magnetic flux path. A detailed schematic of the VRG dynamic test setup is shown in Figure 4-7. The shaker used for the test was a Ling Dynamic Systems (LDS) model V722. The shaker runs off of a closed-loop control cycle with the accelerometer readout fed back to the digital sine controller. Figure 4-8 shows the physical setup of the dynamic tests. The Z-axis control of the shaker, which is the axis running along the center-line axis of the shaker shown in the figure, was used to manually set the minimum air gap by changing the gas pressure below the moving shaker mass. Increasing the pressure below the shaker mass raises the mean position of the shaker element and attached generator piston. This decreased the gap between the piston and stator and subsequently decreased the minimum air gap. The term "minimum air gap" is meant to indicate the distance between the piston and stator air gap surfaces when the piston is at its extreme position closest to the stator. This minimum air gap strongly affects the power output and efficiency of the generator because the magnetic forces are much greater with the same winding current for smaller air gaps compared to larger air gaps. More precise control of the Z-axis control of the shaker and subsequent minimum air gap parameter would have been beneficial but was not possible using the shaker system. To compensate for the poor precision of this variable, data was taken over a range of minimum air gap lengths and data was then compared with other data at similar minimum air gap lengths. The displacements and corresponding air gap lengths were measured using a Philtec D170 fiber-optic displacement sensor. The voltage readout of the displacement sensor was input to a PCIe-6361 National Instruments data-acquisition (DAQ) 108 board. The PCI board was necessary to achieve low latency measurement and control of the generator. The pulse for the gate driver discussed in Section 3.3 was triggered by a rising analog level trigger on the DAQ board. The trigger occurred because the voltage of the fiber-optic sensor increased toward its optical peak as the piston moved closer to the stator. Therefore, the voltage was switched on across the winding when the piston reached a certain displacement on its path moving toward the stator. This trigger point effectively set the turn-on angle, 0n, of the generator The pulse duration was set in the LabVIEW program and precisely controlled by the on-board 100 MHz time-base counter on the DAQ board. The turn-off angle, Gff, is set by the pulse duration. The pulse is sent to the MOSFET gate driver so that the MOSFET is conducting when the pulse is high and an open circuit when the pulse drops to low. Both high and low sides of the circuit where driven with the same pulse signal. Figure 4-8 also shows the method by which the plane formed by the stator poles were made parallel to the plane of the piston laminations. An XY stage and shims placed between two flat aluminum plates above the stator were used to fix four degrees of freedom including the X and Y translation and rotation about the X and Y axes. The rotational degree of freedom about the Z-axis was adjustable by screws on the aluminum component which held the piston laminations. The Z-axis translation was the degree of freedom along which the shaker motion occurred, with the mean position of the plate set by adjusting the mean gas pressure under the shaker plate. Four analog measurements and one digital signal were measured simultaneously using the mixed signal oscilloscope. The digital signal measured was the LabVIEW pulse signal sent to the gate driver on the PCB drive circuit. This signal was used as the trigger for the oscilloscope measurements. The first channel of the oscilloscope was used to measure the voltage readout of the displacement sensor directly so that the displacement amplitude and minimum air gap could be calculated for data analysis. To determine the displacement from the voltage readout, the exact characteristics of the displacement sensor had to be measured. A stainless steel mirror disk was placed on the shaker element as the target for the fiber-optic sensor. The voltage readout of 109 r mr - Physical Connection - - - - - Measurement Set Test Parameter S-r. m Magnetic Unity Gain Differential Force A > Mixed Signal Oscilloscope (Tektronix MS2014B) - --- - -- -- -- -- lIf;m per (1INA105) Ca pacitor Bank (90000 VF) Current Probe (TCP0020) Set Frequency & Amplitude V Wi A W2 AArWwA' Vl LDS Digital Sine Controller (DSC4) Drive Circury Conl (1,2) Cn ,2 W3 Set Displacement Trigger and Pulse Duration i LDS Power Charge Amplifier (PA1000) Accelerometer (PCB J357B01) Gener ator Pisto.n I Fiber-optic Displacement Sensor (D170) LDS Vibrat )r (V722) LDS Field Power Supply (FPS1) Connector Block (NI SCC-68) LabVl EW Software DAQ Board (NI PCIe-6361) Desktop Computer (T5600 Dell) Vibrator Manual Z-axis Control Figure 4-7: This figure details the experimental setup for testing the VRG dynamically. The data from the oscilloscope was saved and exported directly from the oscilloscope to a computer for analysis. The orange lettering refers to the connection points on the PCB schematic shown in Figure 3-17. - XY Stage - Current Probe Shims Z-axis Stage M PCB Fiber-Optic Sensor - Mirror Target Drive Circuitry Stator/Winding Piston M Accelerometer Shaker Plate Figure 4-8: This figure shows the physical setup of the dynamic test used for validation of the VRG. The XY stage and shims were used for aligning the piston and stator. The Z-axis stage was used to calibrate the displacement sensor. 110 the optical peak of the fiber-optic sensor was set to 9 volts. The zero air gap between the piston and stator was set to 8.75 V, which is at the upper end of the linear backslope region of the fiber-optic sensor. The slope of the linear region was measured to be 481.68 mV/mm. The sensor experienced some drift and it was necessary to allow approximately 30 minutes for the sensor to warm-up , and periodic checks were necessary to make sure the optical peak and zero air gap voltages had not shifted so that data was comparable from test to test. This displacement measurement and the corresponding true minimum gap dimension were the greatest source of uncertainty because the drift of the displacement sensor could be up to 50 mV, which is ap- proximately +0.1 mm. This is particularly relevant when comparing experimental data to the model, where the minimum air gap has a significant affect on the power output and efficiency of the generator. The second channel of the oscilloscope was used to measure the voltage across the generator winding terminals. This measurement was a floating voltage with no reference to ground, considering that the applied voltage would switch from positive to negative each cycle. The voltages were applied across a unity gain differential amplifier (INA105), which allowed the voltage difference between the winding terminals to be measured with respect to ground. This was necessary because the 4 channels of the oscilloscope must share the same ground reference. The third channel of the oscilloscope was used to measure the current through the winding using the TCP0020 current probe, and the fourth channel was used to measure the voltage on the bank of capacitors, which were used to store the energy necessary to drive the initial current in the winding. To dissipate the power converted from the mechanical motion to electric power, 5 watt zener diodes were reverse biased from the capacitor bank to ground as shown in Figure 4-9. In this configuration, the zener diodes acted as voltage limiters and dissipated any excess charge driven into the capacitors. A heat sink was manufactured to prevent the diodes from overheating. From this experimental setup, the work output per cycle was determined by calculating the area of the flux linkage-current loop based on the winding applied voltage 111 Gate Driver Wnding Zener 5W a citor Bank Zener 5W Gate Drivere Figure 4-9: The zener diodes were used in this configuration to dissipate the electric power, while maintaining a constant drive voltage. and the current passing through the winding as measured with the current probe. The true power output was calculated based on the winding current and the voltage of the capacitor bank. These measurements were taken for various turn-on and turnoff angles, and frequencies to characterize the power output potential and conversion efficiency of the VRG. These results are presented in the following section. 4.3 Dynamic Experimental Results For the dynamic tests of the VRG, the piston was oscillated at 50 and 60 Hertz with a peak-to-peak amplitude of 5 mm. The power output is significantly lower for these frequencies than the proposed 250 Hertz thermoacoustic engine operating frequency, but the model and VRG performance can still be characterized. Using the trigger voltage and pulse duration as the control variables, the model predicts power and efficiency curves as shown in Figure 4-10. Although the clockwise integration of the flux linkage-current loop produces a negative work output per cycle, the total power output referred to in this thesis is the absolute value of the work output for one cycle multiplied by the operating frequency. This was done to eliminate ambiguity with negative powers. Thus, power output is the generated power that is converted from mechanical oscillations to electric power. The turn-on angle is set by the trigger voltage and maximum/minimum voltages 112 Predicted Efficiency Curves for Fixed Turn-On Angles .I 3. Predicted Power Curves for Fixed Turn-On Angles 5 . S7. Turn-On Angle=2.71 Turn-On Angle=2.54 Turn-On Angle=2.41 Turn-On Angte=2.29 - Turn-On Angle=2.18 Turn-On Angle=2.09 - 0.6 0.5k rad rad rad rad rad rad - 3 _ Turn-On Angle=2.71 Turn-On Angle=2.54 Turn-On Angle=2.41 Turn-On Angle=2.29 Turn-On Angle=2.18 Turn-On Angle=2.09 rad rad rad rad rad rad 2.5 0.4 2 0 w 0.3 0 0.2- 01 0.1 0 1.5 0.5 10 20 30 Pulse Duration (% of Cycle) 40 0 50 10 (a) 30 20 Pulse Duration (% of Cycle) 40 50 (b) Figure 4-10: This figure depicts (a) the predicted efficiency of the VRG operating at 50Hz with a minimum air gap of 0.5 mm and a 10 volt driving voltage. The curves are shown for fixed turn-on angles while varying the pulse duration. (b) Shows the predicted power curve for the same operating conditions. There are clear peaks in both of these plots with the maximum efficiency occurring at shorter pulse durations than the maximum power. for one cycle as determined by the displacement sensor. The turn-on angle (in radians) is then given by GOn = 7r - COS 1 ((VTrigger VMax + VMin 2 2 VMax - VMin (4.4) The pi term comes from the 180 degree shift between the displacement sensor and the minimum gap, where the maximum displacement voltage corresponds to the minimum air gap. As seen in Figure 4-10, there is a local optimum for power and efficiency for each turn-on angle. The shorter the pulse duration and the closer the turn-on occurs to the minimum air gap (Gmin occurs at 0 = r) the less power is lost due to the winding resistance, but this also means the magnetic attractive forces are lower and power output is diminished because there is less time for current to build up in the coil. When the pulse duration is long and the turn-on angle is significantly before the 113 minimum air gap occurs, the winding losses dominate additional power output gains and an overall optimum in power output can be found. This can be seen in part (b) of the figure moving from the peak output power of the purple line (Turn-On Angle = 2.18 radians) to the peak output power of the yellow line (Turn-On Angle = 2.09 radians), where the additional winding losses have caused the maximum output power to decrease even though larger currents (larger magnetic forces) are present. The model does not include the power loss due to the drive circuit components, such as the MOSFETs and diodes, and also neglects the resistance of the wires and PCB traces, with the exception of the copper wire used in the winding itself. The core losses were neglected for the data analysis in this section, but are considered in greater detail in Section 4.3.1. For this reason, to get an accurate comparison of the model to the experiment data, the power output is considered the total power output minus the winding losses, where total power is the area of the flux linkage-current loop (cycle work output) multiplied by the frequency. The predicted power output is the model predicted total power output minus the model predicted winding losses. This method provides a direct comparison of experimental data to model predictions by essentially neglecting losses in the experimental data not accounted for in the model. The efficiency is then defined as the power output divided by the total power output, and the predicted efficiency is the model predicted power output divided by the predicted total power output. When necessary, the actual power dissipated by the zener diodes is specified as true power output, and the true efficiency is the true power output divided by total power output. Shown in Figure 4-11 are a few of the characteristic flux linkage-current loops for the collected data as well as the model predicted loops. These plots were generated by integrating the winding differential voltage to get flux linkage values and the flux linkage and current data were then plotted at each incremental measurement, which traces a clockwise path around the loop for each cycle. Comparison of the model and data was done by specifying identical parameters in the model as those measured in the data. These parameters included the turn-on angle, pulse duration, frequency, applied voltage, and minimum air gap. 114 There are a few possibilities for the deviation between the predicted and actual flux likage-current loops. For the 50 Hertz tests, the model tended to overestimate the maximum flux linkage. This was caused by the capacitor bank shown in Figure 4-9 having a limited capacitance. The voltage drop during the winding charging stage of the cycle shown in Figure 3-18 caused by the finite capacitance led to a smaller maximum flux linkage than predicted by the model. To try and reduce this error, the capacitance of the capacitor bank was increased from 50, 000 to 90, 000 pF between the 50 and 60 Hertz tests. Shown in part (b) of Figure 4-11 is the flux link-current loop for a 60 Hertz test which was driven into saturation. The divergence between the model and the data in this instance is clear in the saturation region, where the inductance of the model is less than the actual inductance. This error is caused by the piecewise linear approach as shown in Figure 4-5, and is an inherent limitation of the model approach used in this thesis. The minimum air gap was the third major contributor to differences between the predicted and actual flux linkage-current paths. This was particularly true for measurements with a short pulse duration because small deviations in the maximum inductance significantly changed the initial trajectory and subsequent path around the flux linkage-current loop. These instances are relatively unimportant because little power can be generated with short pulse durations. The total power, power output, and efficiency were then calculated for each measurement. The results of the 50 Hertz tests for power and efficiency are shown in Figures 4-12 and 4-13 respectively. The data is categorized by the minimum air gap, where the red circles are data points for when the piston at its closest point to the stator during the cycle was less than 0.4 mm from the stator air gap surfaces. The green X's are for minimum air gaps between 0.4 mm and 0.45 mm, the black crosses for minimum air gaps between 0.45 mm and 0.5 mm, the blue stars for minimum air gaps between 0.5 mm and 0.55 mm, the blue squares for minimum air gaps between 0.55 mm and 0.6 mm, and the magenta diamonds for minimum air gaps greater than 0.6 mm. This notation is constant in all of the plots of experimental data on power 115 Predicted vs. Measured Flux Linkage-Current Loop Predicted vs. Measured Flux Linkage-Current Loop 0.045 0.035- 0.04 0.03 0.035 0.025 0.03 --- Predictd 0.025 - - 0.02 X X x x 0.02- 0.015 - 0.01 0.01 0.005 0.005 0 4 00 2 6 4 Current (A) Current (A) (a) (b) 8 10 12 Figure 4-11: This figure shows (a) the measured versus predicted flux linkage-current loops for a 50 Hz test at a turn-on angle of 2.65 radians and a 22.5% duty cycle, and (b) a 60 Hz test driven to saturation with a turn-on angle of 2.1 radians and a 27% duty cycle. and efficiency. Looking at the first plots in Figures 4-12 and 4-13, there are clear bands of power and efficiency with only a few outliers. These bands show the trend that both the power and efficiency increase as the minimum air gap is decreased. This suggests that the minimum air gap has a strong effect on the power output and efficiency of the engine, which makes sense considering the magnetic forces drop with approximately the square of the air gap length. Therefore, if the gap is on average smaller, the generator will convert more of the mechanical oscillation to acoustic power and will do so at smaller winding currents. The effect of the minimum air gap is considered in more detail later in this section. Also shown in these plots are the local maximums in power output and efficiency as predicted and shown in Figure 4-10. The efficiency peak occurs at a pulse duration (duty cycle) shorter than that of the peak power point, which is also expected considering the increase in winding losses when larger magnetic forces are generated. For the first plots in Figures 4-12 and 4-13, the peak power (2.1 W) occurs at a duty cycle of 22.5%, whereas the maximum efficiency (0.677) occurs at a duty cycle 116 50 Hertz Power Output Data Turn On Angle (Approx. 2.8 rad) Local Power Maximum 2. 5 x *-Small 2 0. Gmin x 2 + 2. 5 Turn On Angle (Approx. 2.65 rad) 3 3. 1. 5 1. 5. 0 1 1X 0. 5 0. 5 Large Gmin K,, + 01 -0. 10 15 20 -0. 5 30 25 0 Pulse Duration (% of Cycle) Turn On Angle (Approx. 2.55 rad) 3 2. 5 2 .5 + 3 1. 5 1 .5* . 2 2. + 0 15 20 30 25 20 15 10 Turn On Angle (Approx. 2.2 rad) + * 1.5 1 0. 5 0 0 20 25 + 2 0.5 15 -0.5 30 Mn Ar Gap <0.4mm 0.4mm <Min Air Gap <0.45mm 0.45mm <Min Air Gap <0.5mm 0.5mm -Min Air Gap <0.55mm 0.55mm <Min Air Gap <0.6mm 0.6mm <Min Air Gap . oX 2.5 1 10 0 Turn On Angle (Approx. 2.3 rad) 1. 5 -0. R 0 X Pulse Duration (% of Cycle) 2 0 0 Pulse Duration (% of Cycle) Power Maximum---*- 2. 5, -0 .51 30 25 + 10 3. 0 A- 0 .5 0 EE CL Turn On Angle (Approx. 2.41 rad) 1 - 1 0. 5. -V. 30 25 20 15 10 Pulse Duration (% of Cycle) * L 0 0 I 8 10 15 0 10 0 20 25 30 Pulse Duration (% of Cycle) Pulse Duration (% of Cycle) Figure 4-12: The power output for a range of turn-on angles and pulse durations are shown for the dynamic testing of the VRG at 50 Hertz. The power output is the absolute value of the work per cycle minus the winding losses multiplied by the operating frequency. Data is grouped by minimum air gap to show the correlation between the minimum air gap and the power output. The legend in the bottom right figure applies to all plots. 117 50 Hertz Efficiency Data Turn On Angle (Approx. 2.8 rad) Turn On Angle (Approx. 2.65 rad) Local Efficiency Maximum 0.8 Efficiency Maximum 0.8 0.6 0.6 0.4 E Minimum Air Gap Bands 0.4 0.2 0.2 0 0 Outlier -0.2 -0.2 10 15 20 25 30 10 15 Pulse Duration (% of Cycle) Turn On Angle (Approx. 2.55 rad) 1 20 25 30 Pulse Duration (% of Cycle) Turn On Angle (Approx. 2.41 rad) 1 0.8 0.6 0.6 0.4 0.4 + 0.8 0.2 0.2 0 0 -0.2 10 15 20 41-- 25 0 -0.2 30 15 10 Pulse Duration (%of Cycle) 20 25 30 Pulse Duration (% of Cycle) Turn On Angle (Approx. 2.3 rad) Turn On Angle (Approx. 2.2 rad) 1 1 0.8 0.8 0.6 0.6 0.4 0.4 0.2 0.2 0 0 -0.2 0 O + Gap <0.4mm 0.4m <Min Air Gap '045mm 0.45mm <Min Air Gap '0.5mm 0.5mm -Min Air Gap <055m 0.55mm 'Min Air Gap '06mm 0.6nmm -Min Air Gap Min Air -0.2 10 15 20 25 30 10 Pulse Duration (% of Cycle) 15 20 25 30 Pulse Duration (% of Cycle) Figure 4-13: The efficiency for a range of turn-on angles and pulse durations are shown for the dynamic testing of the VRG at 50 Hertz. The efficiency is the useful work output per cycle over the total cycle work. Data is grouped by minimum air gap to show the correlation between the minimum air gap and the efficiency. The legend in the bottom right figure applies to all plots. 118 of 15%. These local maximums are shown for specific turn on angles in each plot, but overall maximums can be seen as predicted by the model and shown in Figure 4-10. Therefore, the maximum power for the 50 Hertz tests (2.52 W) occurred with a turn-on angle of 2.3 radians and a pulse duration of 27.5% of the cycle. The maximum efficiency point (0.69) occurred with a turn-on angle of 2.65 radians and a pulse duration of 17.5% of the cycle. These points are labeled in the Figures 4-12 and 4-13. The 60 Hertz data shown in Figures 4-14 and 4-15 shows the same trends as the 50 Hertz data but with an increase in both power and efficiency. For an increase in frequency from 50 to 60 Hertz, an increase of greater than 20% in the power output of the engine is expected. The increase is expected to be greater than 20% because the generator losses such as the winding loss and circuit board component losses remain constant while the net power output gain scales linearly with the operating frequency. This is discussed further in Section 4.3.1. Comparing the data, there is approximately a 50-60% increase between the 50 and 60 Hertz operation. For example, comparing point "A" labeled in Figure 4-12 and point "B" labeled in Figure 4-14, the points are at the same turn on angle and have approximately the same duty cycle and minimum air gap. For these two points, the 60 Hertz power output (3.3 W) increased by 51% compared to the 50 Hertz power output (2.18 W). Also of interest is the comparison of the experimental results to the model predictions. To compare the results to the model, the results of data points that fell within 0.075 mm of the 0.5 mm minimum air gap were averaged and individual power and efficiency curves were created for each turn-on angle. Figure 4-16 shows the difference for the 60 Hertz tests between the predicted power output, the power output, and the true power output. These results show qualitative agreement for the peaks in power output. The quantitative agreement is decent between the predicted power output and the power output and suggests that including the drive component loss and a more accurate core loss prediction would predict the true power output with much greater accuracy. The average error between the model predicted and actual power output is 29.7%. This percent error for the generator output power is exaggerated because at the 50 and 60 Hertz frequencies where tests were conducted, the winding 119 60 Hertz Power Output Data Turn On Angle (Approx. 2.75 rad) Turn On Angle (Approx. 2.52 rad) . 6 5. 5 4- 4. 3 0 2 0 3. 0 - 0. 0 2 1 0 0 10 20 15 30 25 6. 6 5. 5. 4. 25 30 35 Pulse Duration (% of Cycle) Turn On Angle (Approx. 2.41 rad) CL 20 15 10 35 Pulse Duration (% of Cycle) Turn On Angle (Approx. 2.3 rad) .5 + B 0 0. 3 [7 0 0 6) , 2. 3 2 0 0~ , 1 0 0 10 15 20 25 30 15 10 35 Turn On Angle (Approx. 2.19 rad) 6. o 5 X + * 4. 30 35 0. 3. 0 2 (D 3t 0.4mm x Mi Air Gap 0.4mm <Min Air Gap <0.45mm 0.45mm <Min Air Gap <0.5mm 0.5mm 'Min Air Gap <0.55mm 0.55mm -Min Air Gap <0.6mm 0.6mm <Min Air Gap 2 - 4) 25 Turn On Angle (Approx. 2.1 rad) 6. 5. 0 20 Pulse Duration (% of Cycle) Pulse Duration (% of Cycle) 0 0 1 0 0 10 15 20 25 30 10 35 15 20 25 30 35 Pulse Duration (% of Cycle) Pulse Duration (% of Cycle) Figure 4-14: The power output for a range of turn-on angle and pulse durations are shown for the dynamic testing of the VRG at 60 Hertz. A comparison of this figure to Figure 4-12 reveals that the power output is greater for the 60 Hertz tests as expected. The data is again separated by the minimum air gap length. 120 60 Hertz Efficiency Data Turn-On Angle (Approx. 2.52 rad) Turn-On Angle (Approx. 2.75 rad) 0.8 0.8 0.6 r- 0.6 0.4 0.4 0.2 0.2 0 0 -0.2 10 15 20 25 -0.2 30 S+ 15 10 1 0.8 0.8 0.6 0.6 0.4 0.4 0.2 0.2 0 0 -0.2 10 10 15 15 20 20 25 25 30 Turn-On Angle (Approx. 2.3 rad) Turn-On Angle (Approx. 2.41 rad) 1 25 20 Pulse Duration (% of Cycle) Pulse Duration (% of Cycle) -0.2 30 -~ 15 10 25 20 30 Pulse Duration (% of Cycle) Pulse Duration (% of Cycle) Turn-On Angle (Approx. 2.1 rad) Turn-On Angle (Approx. 2.19 rad) 1 0.8 0.8 0.6 0.6 - 1 O 0.2 0.2 X + * E 0 0 -0.2 $1 0.4 10 15 20 25 -0.2 30 10 + I 0.4 Min Air Gap <0.4mm 0.4mm <Min Air Gap <0.45mm 0.45mm <PMn Air Gap <0.5mm 0.5mm <Mn Air Gap <0.55mm 0.55mm <Min Air Gap <0.6mm 0.6mm <Min Air Gap 15 20 25 30 Pulse Duration (% of Cycle) Pulse Duration (% of Cycle) Figure 4-15: The efficiency for a range of turn-on angle and pulse durations are shown for the dynamic testing of the VRG at 60 Hertz. Comparing the data in this figure to Figure 4-13 reveals that for a given turn-on angle, the efficiency is slightly greater for the 60 Hertz operation than the 50 Hertz operation. This is also expected and discussed further in Section 4.3.1 121 60 Hertz Power Output Comparison Turn On Angle (Approx. 2.75 rad) 6 6 5. 5 4. 4 0. 3 3 0 2. 0 0- Turn On Angle (Approx. 2.52 rad) 2 a. 1 0 0 5 10 20 15 25 10 15 Pulse Duration (% of Cycle) Turn On Angle (Approx. 2.41 rad) 6 6 5. Turn On Angle (Approx. 2.3 rad) - 4 3. 0 2 a0 2 0 16 1 3 0 18 20 22 24 26 28 15 20 Pulse Duration (% of Cycle) 6 5 30 25 5 4 0 20 Pulse Duration (% of Cycle) 30 25 Pulse Duration (% of Cycle) Turn On Angle (Approx. 2.19 rad) Turn On Angle (Approx. 2.1 rad) 6 Power Output 5 Predicted Power Output True Power Output 4. 3 0. 2 0 2 3 n 1 1 10 0 - 0 15 20 25 30 20 Pulse Duration (% of Cycle) 22 24 26 28 30 32 Pulse Duration (% of Cycle) Figure 4-16: In this figure, the data for each pulse duration is averaged to produce experimental results plots for power output. The same is done for the model and true power output, which includes the core and inverter losses not included in the model. The model shows good qualitative accuracy to the experimental data and reasonable quantitative accuracy for the power output. 122 60 Hertz Efficiency Comparison Turn-On Angle (Approx. 2.75 rad) 0 .0 0.8 0 .6. 0.6 0 .4 W. 0 .2 Turn-On Angle (Approx. 2.52 rad) 0.4 0.2 01 5 20 15 10 1 0) Turn-On Angle (Approx. 2.41 rad) 0 .0 .4- C.) II 0 .40..2 0..20 16 Turn-On Angle (Approx. 2.3 rad) 0 .6- 0..60 20 18 22 26 24 0 15 23 30 25 20 Pulse Duration (% of Cycle) Pulse Duration (% of Cycle) 0 .8 30 Pulse Duration (% of Cycle) Pulse Duration (% of Cycle) 0..8 25 25 20 20 15 15 0L 25 Turn-On Angle (Approx. 2.1 rad) Turn-On Angle (Approx. 2.19 rad) A Efficiency Predicted Efficiency True Efficiency 0.6 . 0 .6 C, 0 .4 0.4 0 .2 0.2 '0 n 10 15 20 25 220 30 22 24 26 28 30 32 Pulse Duration (% of Cycle) Pulse Duration (% of Cycle) Figure 4-17: In this figure, the data for each pulse duration is averaged to produce experimental results plot for the efficiency. The same is done for the model and true efficiency of the VRG, which includes the core and inverter losses not included in the model. The model shows excellent qualitative and quantitative accuracy for efficiency, suggesting that the model predicts the current waveforms quite accurately for a given applied voltage. 123 True Power Output vs. Minimum Air Gap True Efficiency vs. Minimum Air Gap 0 0 x( 5- + * 0 0 0 Min Air Gap <0.4mm 04mm <Mln AirCGa 6m 0.45mm <MinAirGap<0.5mm 0.5mm <Min Air Gap <0.55mm 0.55mm <M nArGap <0.6mm 0.6mmn<MinAir Gap . . ' x 0.5 0 * 0 > 0 MmA i AGap <0.4mm 04mm <Mi Air Gap <0.45mm . m n r ap . mm 0.5mm <MinAirGap<0.55mm 0.55mm <Mn Air Gap <0.6mm 0.6mm <Min Air Gap 0.4* 4- X x - *+ w 0.3 + ++ ft+ 2 0.2- 2 1 0 E * 3 0.1 0.2 0 0.4 0.6 Minimum Air Gap (mm) 0.8 1 01 0 0.2 0.4 0.6 Minimum Air Gap (mm) (a) 0.8 1 (b) Figure 4-18: (a) The true power output of the engine is shown as a function of the minimum air gap for a turn-on angle of approximately 2.3 radians and a 27% duty cycle. The colors are shown so that the data is comparable to Figures 4-14 and 4-15. The maximum true power output (5.74 Watts) occurs at the smallest minimum air gap (Gmin = 0.325 mm). (b) Shows the true efficiency of the generator as a function of minimum air gap spacing for the same tests as in (a). There is a very strong correlation between the minimum air gap and both power and efficiency. The maximum true efficiency (55.8%) also occurs at the smallest Gmin. losses are nearly as large or larger than the actual power output. This means the power output measurement is very sensitive to winding losses and the accuracy of data measurement. The same comparison of the efficiency between the model and experimental data was done and shown in Figure 4-17. The efficiency predicted by the model and the experimental efficiency are qualitatively very similar. The average difference between the measured and predicted efficiency is 4.6%. This quantitative accuracy suggests that the model predicts the current waveform quite accurately. The efficiency including the inverter losses and core losses is qualitatively similar, but again the difference is exaggerated because of the lower operating frequency at which the testing was done. To examine more closely the effect of the minimum air gap, a set of data points 124 was collected at a range of minimum gap lengths while maintaining the same pulse duration and trigger voltage. The trigger voltage is referenced to the displacement sensor, and therefore, the change in minimum gap spacing results in a slight shift of the turn-on angle. This is seen by Equation 4.4, where VMa, and VMin shift up or down slightly when the piston on average is closer or further away from the stator resulting in smaller or larger minimum air gaps respectively. However, for a shift of 0.2 mm the change in turn-on angle is about 0.1 radians for the 5 mm piston displacement. Therefore, the data is treated as essentially a constant turn-on angle. The data for the various minimum air gap spacings was taken at 60 Hertz, with a trigger voltage of 8.1 volts, and a pulse duration of 4.5 milliseconds. This corresponds to a turn-on angle of approximately 2.3 radians with a 27% duty cycle. This data is presented in Figure 4-18. This operating point was chosen because it was a clear local maximum in power as seen in Figure 4-14. Therefore, for this axial gap design to be effective, the minimum air gap, which results in the maximum inductance, should be as small as possible. For this generator, the maximum true power output generated by the 5 mm piston displacement was 5.74 Watts with a true efficiency of 55.8%. The model underestimates this operating point by 1.5 watts, which suggests that the maximum inductance and saturation characteristics of the piecewise linear model do not exactly align with the true generator characteristics. This would lead to an underestimate in the total work per cycle and overestimate the currents driven through the winding. 4.3.1 Frequency Scaling and Losses The proposed operating frequency of the generator is approximately 4-5 times the frequencies used for model validation and testing of the VRG system. Given the operating frequency of the engine, w, the power dissipated by eddy current losses is proportional to w 2 , the power output is proportional to w, and the winding losses are not affected by frequency [30]. Qualitatively this can be described as in Figure 4-19, where the plotted variables are the fraction of power dissipated in the winding, P., 125 k x Figure 4-19: This figure depicts qualitatively the fractional core losses and winding losses to the power output as a function of frequency. This is particularly important considering the scaling of the engine to higher frequencies where eddy current losses become more significant. Figure adapted from [30]. to power output, P, or Xw (4.5) - PO and the fraction of power dissipated by core losses, P, to power output, P., or Pe Xe = Pe .(4.6) PO' Therefore, the minimum fraction of power dissipated to power output occurs when the power dissipated through winding and core loss are equal [30], given by X = Pe+Pw PO (4.7) The core losses for the M19, 29-gage steel used in the generator are currently being characterized, but the testing is not complete at the time of the writing of this thesis. However, a good approximation can be made based on published data of M19 electric steel. Considering the VRG system designed, the core loss as a function of frequency is shown in Table 4.1. The core loss data comes from data supplied by Polaris Laser Laminations in reference [21]. The core loss was estimated by assuming a saturation of 1.5 Tesla in the piston laminations and 1 Tesla in the stator laminations, given that 126 Table 4.1: Estimated VRG Core Loss Core Loss (per phase) Frequency (Hz) 50 60 100 150 200 300 Piston (W) 0.02 0.07 0.13 0.22 0.34 0.61 Stator (W) 0.05 0.06 0.12 0.20 0.31 0.55 Total (W) 0.07 0.13 0.25 0.42 0.65 1.17 the cross-sectional area of the stator is approximately 50% greater than the piston. The core loss is given in terms of watts per kilogram, so that core loss could be approximated based on the known mass of the steel laminations (181.6 gram stator, 42.8 gram piston). The core loss of the published data was subsequently multiplied by one half because the minor hysteresis loop traveled is roughly half the full hysteresis loop as discussed in Section 3.2.3. Given that the winding loss for the 50 and 60 Hertz tests was approximately 3-4 watts when driven near the the maximums of power output, the eddy current loss does not become as significant as the winding loss until frequencies significantly greater than the proposed 250 Hertz operating point even when the steel is driven into saturation. For this analysis, it was determined that the model underestimates the magnitude of the eddy current loss by up to a factor of 2 using Equation (3.29). Nonetheless, by scaling the results of the dynamic testing and increasing the effect of eddy current by a factor of two, the model predicts the VRG is easily capable of the desired 25 watts per phase power output with a conversion efficiency in the range of 80-85%, with potential for greater efficiency if less than 0.5 mm can be held for the minimum air gap. The experimental data can also be scaled to higher frequency with the known scaling proportion of losses and power output. Figure 4-20 shows the true power output and true efficiency by scaling the 60 Hertz data to 250 Hertz. The data shown is for a turn-on angle of 2.3 radians. The data was scaled by increasing the total power in proportion to the frequency, increasing the core losses in accordance with 127 Scaled 250 Hz True Power Output . . . . . Scaled 250 Hz True Efficiency 0.02 . n. 70 0.8F 60 - 0.78P- Tum-On Angle = 2.3 radl 50 0 0 C 40 0.76 0 4E ---- Tum-On Angle = 2.3 rad 0 2 0.7220 0.7- 10 0.18 0.2 0.28 0.26 0.24 0.22 Pulse Duration (% of Cycle) 0.18 0.3 0.2 0.22 0.24 0.26 0.28 0.3 Pulse Duration (% of Cycle) (b) (a) Figure 4-20: (a) The 60 Hertz data (0o0 = 2.3 rad) for true power output is shown scaled to the 250 Hertz operating point by appropriately scaling the power output and generator losses. (b) The true efficiency is shown for the same operating points as in (a). These figures show that the generator is capable of producing the desired 50 Watt output at the higher operating frequency. Additional power output may be possible by increasing the duty cycle (pulse duration), but the generator efficiency begins to drop off quickly. Table 4.1, and assuming all other losses are constant. The data was also multiplied by two to give true power output when operating with both phases of the generator. The scaled data shows that the VRG will be able to generate greater than 50 Watts of electric power with an efficiency of 80-85%. If the peak operating point for the generator shown in Figure 4-18 (Gmrn = 3.2 mm) is scaled to 250 Hertz, the true power output of the generator is 74.8 Watts with a true conversion efficiency of 87.3%. This again shows the importance of the minimum air gap and predicts that the generator can be highly efficient for convert- ing high frequency, small amplitude oscillations if the minimum air gap is precisely controlled. 128 4.4 Chapter Summary The VRG system was tested both statically and dynamically to characterize the power output of the system and to determine the accuracy of the model by which the VRG was designed. The results of the static testing may be used to create a more accurate model using similar methods as [29]. The qualitative predictions of the model matched very well with the collected data. The model's quantitative predictions were less accurate, which is attributed to the inaccuracy of the measurement of the minimum air gap, the drop in applied voltage across the capacitor during the charging of the winding, and to the model's piecewise-linear approximation, which is particularly relevant when the steel saturates. The maximum output of the VRG at 60 Hz was measured at 5.74 watts with an efficiency of 55.8%. From the experimental data, it is concluded that the VRG system will be able to produce the desired 50 watt output from the 250 Hertz thermoacoustic engine oscillations with an efficiency of approximately 80-85%. 129 Chapter 5 Conclusions and Future Work 5.1 Conclusions This work has covered a full thermal-to-electric power system concept based on thermoacoustic technology. The three separate parts of the design are the thermoacoustic engine in which thermal energy is converted into mechanical pressure oscillations, an electroacoustic transducer, and a self-pumping gas bearing system. The engine was designed to produce 50-100 Watts of electric power from a robust, portable system to serve as a remote power source for soldiers. The thermoacoustic system was designed to operate as a double Helmholtz-like resonator with a mechanically resonant piston as the inductance element of the resonator. This allows for a compact design and the piston to act as an electroacoustic transducer that is strongly coupled to both the electromagnetic system as well as the acoustic system. The thermoacoustic engine was designed and optimized using DeltaEC software. The electroacoustic transducer based on variable reluctance principles was designed, fabricated, and tested. The cross-shaped design of the transducer allowed the suppression of eddy currents and allowed for rotational stability of the piston assembly. The impetus for the new linear alternator design was the necessity of coupling the transducer with higher-frequency and smaller displacement amplitudes than typical linear alternators are optimized for. The design was evaluated and optimized for the 130 thermoacoustic system based on a piecewise linear saturation computational model of the VRG system. The model and VRG system were then validated based on static and dynamic tests. The static tests led to a reevaluation of fringing effects in the generator. The dynamic tests revealed excellent qualitative agreement between the model predictions and the experimental data. The quantitative predictions were accurate within approximately 15% for the high power tests of the VRG system. The minimum air gap length was shown to have a significant effect on both the power output and efficiency of the generator. This led to the conclusion that for the VRG system with an axial air gap to be effective, the minimum air gap should be kept as small as possible and must be less than approximately 0.6 mm to be a realistic option for power transduction. At its peak operating point, the VRG system generated 5.74 Watts of electric power with an efficiency of 55.8% at a frequency of 60 Hertz. Given the scaling of power output to losses of the VRG with increasing operating frequency to 250 Hertz, the VRG will be capable of producing at least 50 Watts of electric power with an efficiency of 80-85%. Scaling the peak operating point to 250 Hertz suggested a true power of 74.8 Watts is possible with a conversion efficiency of 87.3% if the minimum air gap can be held to 0.3 mm, and potentially better operation if the minimum gap can be kept smaller. The gas bearing system designed for the VRG is beneficial for both engine operating lifetimes and removing suspension resonance and displacement limitations. The bearing system proposed relies on the motion of the piston to selectively open and close ports with a specific timing to produce high and low pressure reservoirs. In this configuration, a passive bearing system is produced that keeps the piston centered. The primary limitation of the bearing system is the necessity for extremely tight tolerances of machined surfaces to minimize the gap between the piston and cylinder wall. This bearing system as well as other components of the engine have been left as future work for the project, which is discussed in the following section. 131 5.2 Future Work The compilation of the thermoacoustic engine, linearly-acting variable-reluctance generator, and the gas bearing system presented in this thesis is left as future work for the project, as well as the other recommendations described in this section. 5.2.1 Modeling There are a few improvements to the models that should be completed to better predict the performance of the full thermoacoustic generator system. For the thermoacoustic portion of the engine, incorporating the nonlinear transducer with the DeltaEC code or other thermoacoustic coupling analysis should be completed. This would present more accurately the magnetic forces and there effect on the motion of the piston given the relative phasing of the pressure oscillations. The DeltaEC model should also be updated with specific component parameters as a more complete design of the stack, cold heat exchanger, and hot heat exchanger are completed. Additionally, the piecewise-linear approach for modeling the VRG can be improved by using the flux linkage-current data produced for the generator. This would eliminate errors in the projected performance particularly when operating the generator in high power applications where the steel becomes saturated. This would also present a more effective method for determining the optimum turn-on and turn-off angles for the VRG system. For the bearing system, additional modeling should be done including better approximations of the flow resistances, the inclusion of gas compressibility, and potentially turbulence in orifice type restrictions. Additional consideration should be given to the thrust bearing application of the bearing system so that the piston operates on average in exactly the center position (axially) of the generator. Finally, an analysis of bearing start-up and liftoff are important modeling considerations for the application of the gas bearing system. 132 5.2.2 System and Fabrication Improvements The thermoacoustic system is capable of operating from any quality heat source. How the thermoacoustic hot heat exchanger interfaces with the heat source should be considered, whether that heat source is a butane heater or a solar type application. The vision of this researcher is to produce a thermoacoustic engine with an interface similar to the commercial Jetboil system but with the aim of producing electric power instead of cooking food. Additionally, the cold heat exchanger was modeled as operating at 340 Kelvin to account for an air cooled cold heat exchange system because a more effective water cooled system is not a practical option for a portable system. Further design of the cold heat exchanger is required. The control of the VRG system is an important consideration of the final system. Sensorless control of the VRG (using the coils to determine the piston of the piston) should be considered to remove the required displacement sensor that was used in the experimental testing of the VRG system. Additionally, using the timing of the VRG turn-on and turn-off angles to precisely control the location of the piston should be considered. This piston control is a secondary option to using the bearing system in a thrust bearing type application. This engine should be built and tested, but as a long term improvement on the thermoacoustic system, the standing wave bounce volumes could be operated as torus type traveling wave bounce volumes as in the Swift traveling wave engine. This would be an important improvement if the efficiency of the system becomes a more important consideration. 133 Appendix A Flux Tube Analysis The estimation of fringing fluxes was necessary in the design of the VRG system because of the large air gaps where fringing effects significantly increase the minimum inductance. Figure 3-14 is repeated here in Figure A-i for clarity when describing in more detailed the flux tube modeling done for the generator. The fringing types are divided into surface-to-surface simple shaped volumes, edge-to-edge simple shaped volumes, and corner-to-corner simple shaped volumes. Surface-to-surface are referenced by the two letters defining the two surfaces, edges are defined by the surfaces which meet to form the edge, and corners are defined by the three surfaces which meet to form the corner. For derivations of these equations and for more description on the specific simple shaped volumes see [18]. The following equations were used to calculate the flux path permeances. For clarity, there is a "G" that corresponds to a surface in the figure, but in the equations "G" refers to the air gap length. Outer Air Gap PA=>G = PAB=>GC = G 2(0.26Iu(lo + z)) (A.1) (A.2) where the multiplication by 2 corresponds to the 2 sides of the stator. Numbers in front of the equation will continue to reference the number of times that geometry 134 F 0 A H I 0 D Figure A-1: This figure shows two of the simple-shaped volumes used for the flux tube analysis. The half cylinder represents the flux from edge AB to edge GF, and the half annulus represents the flux from surface B to surface F. occurs for one stator piece-piston piece pair. 26 w PAD=>EG =0. PB=>F =2 PD=E = + G )) (A.4) G1)) (A.5) (L/r n (I + PABD=>EFG = PBD=>EF = 0 PAC=G PC=>G 2hp z) 1 (l+ w (A-3) 4(.077pG) (A.6) 2 (A.7) (4) . 1 135 5 2 tpw 1P (A.8) (A.9) Inner Air Gap (A.10) PI=>- = PH=G = 0.52pw +L w1 PH=>-G PBHI=>FG= (A.11) -A12) 2(0.077 1 tG) (A.13) (Lv) -2 (A.14) PBH=>F Leakage Permeance PL = p(A.15) 1w Each of these permeances were present on each of the 4 stators, so the permeances here were multiplied by 4 to calculate the fringing and leakage fluxes for one phase of the full generator. 136 Appendix B Model Code and Experimental Data The DeltaEC model used for analysis and optimization of the thermoacoustic engine is provided here. Three-Inch Thermoacoustic Engine TITLE 21-Feb-2014 with DeltaEC version 6.3b11.12!under win32, !CreatedQ03:47:33 using Win 6.1.7601 (Service Pack 1) under Python DeltaEC. !--------------------------------- 0 --------------------------------Initial BEGIN 3.0000E+06 a Mean P Pa 253.51 b Freq Hz 659.54 c TBeg K G Ip1 Pa G 1.5727E+05 d 0.0000 a Ph(p) 0.0000 : IUI dog m^3/s 0.0000 g Ph(U) helium G dog Gas type !--------------------------------- 1 --------------------------------SURFACE Hot End 4.5600E-03 a Area 1.5727E+05 A IpI m-2 0.0000 B Ph(p) 1.3099E-05 C 180.00 IUI D Ph(U) 0.0000 E Htot ideal Solid type -1.0301 ?--------------------------------- 2 deg m-3/s deg W W --------------------------------- Hot Duct DUCT sameas la a Area 0.23938 b Perim 2.5000E-02 c Length m-2 Mstr 1.5722E+06 m a 2a 2.2548E-04 B Ph(p) deg 5.7284E-03 C 1UI m^3/s -90.303 ideal 3 A IpI D Ph(U) Pa deg 0.0000 E Htot W F Edot W -2.382 Solid type !--------------------------------HX F Edot Pa --------------------------------- Hot HX 3.80OOE-03 a Area m-2 1.5717E+05 A IpI Pa 137 0.6700 b GasA/A 5.0000E-03 c Length m 5.2540E-04 d 194.12 e m 580.00 yO HeatIn W G Solid type !---------------------------- ----- STKSLAB Stack sameas 4a a Area 0.8200 4 m-2 D Ph(U) 194.12 E W F Edot W 659.54 G GasT K 668.11 H SolidT K --------------------------------- 1.1175E-02 C IUI 5.0000E-04 e Lplate mn Solid type Stack sameas 4a a Area sameas 6b b GasA/A sameas 6c c Length mn sameas 6d d yO m-2 m^3/s D Ph(U) dog 194.12 E Ntot W F Edot W 659.54 G TBeg K 446.22 H TEnd K 1.5578E+05 A Ipi 0.31691 B Ph(p) 1.6047E-02 C IUI m 5.0000E-04 e Lplate m Solid type *--------------------------------- 6 Pa deg m-3/s -88.179 D Ph(U) dog 194.12 E Htot W F Edot W 446.22 G TBeg K 346.98 H TEnd K 32.812 stainless deg --------------------------------- !---------------------------------5 STKSLAB Pa -88.824 15.785 stainless dog Htot 0.14281 B Ph(p) m m 1.4000E-04 d yO m3/s -90.887 1.5672E+05 A Ip1 b GasA/A 3.0000E-02 c Length deg 6.4358E-03 C IUI -7.852 f SolidT K ideal 3.0308E-03 B Ph(p) --------------------------------- Ambient HX HX sameas 4a a Area 0.2530 m-2 1.5504E+05 A Ip1 b GasA/A 0.34036 B Ph(p) 6.0000E-03 c Length mn 4.0600E-04 d yO e -169.12 340.00 -88.288 Is HeatIn W f SolidT K ideal 1.6355E-02 C IUI m^3/s dog 25.000 E -13H 30.341 F Edot W 346.98 G GasT K 340.00 H SolidT K !--------------------------------- 7 Htot deg G Solid type DUCT D Ph(U) Pa W --------------------------------- Ambient Duct samea. la a Area 0.23938 b Perim m-2 m 1.5484E+05 A Ip1 0.33884 B Ph(p) 8.0000E-03 c Length m 1.8159E-02 C IUI Solid type ideal !--------------------------------RPN Pa dog m'3/s -88.435 D Ph(U) deg 25.000 E Htot w 30.096 F Edot w 8 Piston Disp 2.5000E-03 a G or T -16A A Piston 2.5000E-03 / Ul mag la / v I--------------------------------IESPEAKER sameas la a Area 0.2500 b R 0.0000 c L 18.000 0.2000 0.0000 m-2 1.5484E+05 A Ipi ohms -179.66 H 1.8159E-02 C IUI B Ph(p) Pa deg m'3/s d BLProd T-m -88.435 D Ph(U) e M kg -25.00 E Htot w f N/m -30.096 F Edot -50.00 G WorkIn w W K 0.1000 g Rm N-s/m 8.1227 h III A i deg -10.00 9 Change Me Ph(I) G 71.300 8.1227 99.943 H Volts I Amps dog V A 3 Ph(V/I) deg 138 3.0968E+05 K IPxI -179.66 Solid type ideal 10 RPN Pa L Ph(Px) deg --------------------------------- ChangeMe -50.00 A ChngeMe -50.00 -18A a G or T 17G !--------------------------------- 11 --------------------------------- Change Me DUCT sameas la a Area 0.23938 b Perim sameas m-2 Mstr 1.5504E+05 A a 19a -179.66 14c c Length a 1.6355E-02 C lUl Solid type ideal IpI B Ph(p) i--------------------------------- Pa deg m^3/s -88.288 D Ph(U) deg -25.00 E Htot w -30.341 F Edot w 12 Ambient2 HX sameas 1.5578E+05 A IpI Pa sameas 13b b GasA/A -179.68 deg sameas sameas 13c c Length a 1.6047E-02 C IUI m-2 4a a Area m 13d d yO -88.179 -169.12 e HeatIn W G 340.00 f SolidT K -20H -194.12 -32.812 Solid type ideal !--------------------------------- B Ph(p) D Ph(U) F Edot 346.98 G GasT 340.00 H SolidT K K 13 --------------------------------- Me Change sameas 4a a Area sameas sameas 6b b GazA/A -179.86 6c c Length a 1.1175E-02 C 1Ut m-2 1.5672E+05 A Ip1 M 5.0000E-04 e Lplate m -88.824 6d d yO !--------------------------------- B Ph(p) Pa deg m^3/s D Ph(U) deg E Htot v -15.785 F Edot w 346.98 G TBeg K 446.22 H TEnd K -194.12 Solid type stainless 14 --------------------------------- Change He STKSLAB m-2 1.5717E+05 A IpI sameas 4a a Area sameas sneas sameas 6b b GasA/A -180.0 6c c Length a 6.4358E-03 C 1UI m 21d d yO -90.887 6.0000E-04 e Lplate m -194.12 Solid type stainless I--------------------------------- 15 B Ph(p) Pa deg m^3/s D Ph(U) deg E Htot w F Edot w 446.22 G TBeg K 659.54 H TEnd K 7.8520 HX deg E Htot STKSLAB sareas m-3/s --------------------------------- Hot2 m-2 Ip1 sameas 4a a Area sameas 4b b GasA/A -180.0 sameas 4c c Length a 5.7284E-03 C 1U1 samoas 4d d yO sameas sameas 4e e HeatIn W ideal 1.5722E+05 A M -90.303 4f f SolidT K Solid type Change sameas Pa deg M^3/s deg 2.3820 F Edot w w 659.54 G GasT K 668.11 H SolidT K 16 Me Ia a Area 0.23938 b Perim sameas D Ph(U) 9.1518E-12 E Htot !--------------------------------DUCT B Ph(p) m^2 m 2c c Length a Mstr 25a 1.5727E+05 A 1p1 180.00 B Ph(p) 1.3099E-05 C IUI -180.0 D Ph(U) 9.1518E-12 E Htot Pa deg mr3/s deg w 139 ideal Solid type 1.0301 F Edot SURFACE Change sameas la a Area Me m-2 1.5727E+05 A Ip1 180.00 B Ph(p) 9.3406E-15 C -81.797 ideal Solid type !--------------------------------HARDEND W 17 --------------------------------- I--------------------------------- IUI D Ph(U) Pa deg mi3/s deg 9.1518E-12 E Htot W -1.0480E-10 F Edot W 18 --------------------------------- target this to seal the end 0.0000 a R(1/z) -27G 0.0000 b I(1/z) -27H 0.0000 c Htot W 1.5727E+05 A Ip1 180.00 -27E B 9.3406E-15 C -81.797 Pa Ph(p) deg IUI m^3/s D Ph(U) deg 9.1518E-12 E Htot W -1.0480E-10 F Edot W -6.1486E-15 G R(1/z) 4.2653E-i4 H I(1/z) I The restart information below was generated by a previous run I and will be used by DeltaEC the next time it opens this file. guesez Ob Oc Od 4e 13e xprecn 2.0190E-03 -5.8246E-04 targe 13f mstr-slave 16a 18a 20f 27a 17h 20e 3.9218 27b 9.2722E-04 -9.2722E-04 -2.8523E-06 -9.2722E-04 27c 3 2 -2 19 -2 25 -2 I Plot start, end, and step values. Outer Loop: May be edited if you wish. I Inner Loop 140 Appendix C Matlab Code for VRG The Matlab model of the VRG transducer is provided here. Variable Initialization clear all %Circuit Time .Number of divisions per cycle n-10000; Angle theta-linspace(0,2*pi,n)'; XPhase Freq-250; %Operating frequency omega-Freq*2*pi; XDisplacement, Angular frequency %Capacitor bank capacitance C-50000; Clearance, Position, Velocity Disp-0.0060; Peak-to-peak displacement amplitude Clear-0.0006; %Minimum Air Gap x-Disp/2*cos(theta)+Disp/2+Clear; Y.Air gap length (all lengths in meters), "G" used in thesis %VRG Design Constants ItM-.0254; %Inch to Meter Conversion muo-4*pi*10^-7; XPermeability mur-3000; of free space Approx. steel relative permeability mu-muo*mur; %Eddy Current Parameters rhos-5*10-7; tlam-3.46*10^-4; %Steel resistivity (Ohm m) Laination Thicknes %Component Dimensions (Same notation as in thesis) li-.31*ItM; z-. 1*ItM; w-2*(li-z); Ag-4*li2-4*z-2; lo-Ag/(4*v); lp-sqrt(.0367^2-v^2); lv-lp-(li+.021*ItM)-lo-.0018923; Y.Constant used for winding clearance hc-.25*ItM; hv-.626*ItM; hp-.1 *ItM; As-4*hc*w; Ap-4*hp*w; A-[Ap, As, Ag]; Amin-min(A); 141 Leakage and Fringing Permeances % Leakage and Fringing Permeances CF-1.4; ZFringing correction Factor Pi-4muo*lo*w./x*CF; P2-2*4*.26*muo*lo*CF; P3-4*.26*muo*w*CF; P4-2*4*muo*lo/pi*log(1+2*hp. /x)*CF; P5-4uo*w/pi*log(1+2*hp. /x)*CF; P6-4*4*.077*no.*x*CF; P7-4*muo*hp/4*CF; P8-4*.52*muo*w*CF; P9-4*2*muo*w/pi*log(i+hp./x)*CF; PL-4*muo*hw*w/lw; P11-muo*Ag./x*CF; P12-4*.52*muo*w*CF; P13-4*2*muo*w/pi*log(i+hp./x)*CF; P14-2*4*.077*muo.*x*CF; P15-2*4*muo*hp/4*CF; %Reluctances Ri-2*hw/(mu*Ag); R2-lp/(mu*hp*w); R3-R1; R4-lp/(mu*hc*w); RL-2/PL; Rf-1i./(Pi2+Pi3+P14+Pi5); Rg1-1./P11; Rf2-i. /(P2+P3+P4+P5+P6+P7+P8+P9); Rg2-1./P1; Rp-./(/RL+i./(R2+1./(i./Rfi+i./Rgi)+i./(i./Rf2+1./Rg2))); Rtot-R1+R3+R4+Rp; Wire Resistance XWire Resistance/Number of turns based on wire gauge N-148; Number rhow-16.8*10^-9; Winding resistivity of turns Ff-.pi/4; %Wire packing factor Aw-(lv-.032*ItM)*(hw-.032*ItM); XConstant Lw-8*(li+lw/2); XWire AWGA-Aw*Ff/N; %Wire area AWG-sqrt(4/pi*AWGA); Y.American Wire Gauge R-rhow*Lw*N.^2/(Ff*Aw); %Wire resistance used for winding clearance length Material Saturation Characteristics UInductance L-N^2. /(Rtot) ; UInductance (H) Ls-10*pi*10^-4; XConstant saturated inductance %Gap Saturation Current Bs-1.6; %Assumed saturation flux density (T) Is-Bs*Amin*Rtot/N; %Saturation Current (A) Drive Circuitry Variables Batt-50; ThetaOn-2.6; %Capacitor Bank Voltage Turn on angle LVTimeOn-.0009; %Pulse duration ThetaOff-LVTimeOn*omega+ThetaOn; Y.Turn off angle 142 Initializing Cyclic Current and Flux Linkage Profiles %Calculating current vaveform Isat-ones(n, 1); I-ones(n,i); Lambda-ones(n,i); Vb-ones(n, 1); V-ones(n,i); Bp-ones(n, 1); Bc-ones(n,1); Theta-ones(n,i); dBpdth-ones(n, 1); dBcdth-ones(n,1); dLdxcur-ones(n,1); L-ones(n,i); vel-ones(n,i); F-ones(n,i); Test-ones (n,1); %Loop to determine initial current waveform/ flux linkage for m-i:n angl 0 Theta(m)-(m-1)/n*2*pi; XPhase xp-x(m); Y.Air gap/pi ston displacement vel(m)--Disp/2*omega*sin(Theta(m)); %Piston vel oicty % Leakage and Fringing Permeances Pi-4*muo*lo*w/xp*CF; P4-2*4*muo*lo/pi*log(1+2*hp/xp)*CF; P5-4*muo*w/pi*log(1+2*hp/xp)*CF; P6-44*.077*muo*xp*CF; P9-4*2*muo*w/pi*log(i+hp/xp)*CF; P1i-muo*Ag/xp*CF; P13-4*2*muo*w/pi*log(1+hp/xp)*CF; P14-2*4*.077*muo*xp*CF; %Reluctances Rf 1-1/(P12+P13+P14+Pi6); Rg1-1/P11; Rf2-1/(P2+P3+P4+P5+P6+P7+P8+P9); Rg2-1/Pi; Rp-1/(i/RL+1/(R2+1/(i/Rfi+/Rg)+1/(1/Rf2+1/Rg2))); Rtot-R+R3+R4+Rp; L(m)-N-2/Rtot; Isat (m)-Bs*Ap*Rtot/N; if Theta(m)<-ThetaOn Lambda(m)-0; I(m)-0; Vb(m)-0; V(m)-0; else if (Theta(m)>ThetaOn U Theta(m)<-Thetaff) Vb(m)-Batt; Lambda(m) -Vb (i) /omega* (Theta(m) -ThetaOn); I(m)-Lambda(m)/L(m); V(m)-Vb(m)-I(m)*Ract; if V(m)<O Lambda(m)-Lambda(m-1); I(m)-Vb(m)/Ract; end else if (Theta(m)>ThetaOff U Lambda(m-1)>O) Vb(m)--Batt; Lambda(m) -Batt/omega* (ThetaOff-ThetaOn) +Vb(m) /omega* (Theta(m) -ThetaOff); I(m)-Lambda(m)/L(m); V(m)-Vb(m)-I(m)*Ract; if V(m)>O Lambda(m)-Lambda(m-1); 143 I(m)--V(bm)/Ract; end else Lambda(m)-Lambda(m-1); I(m)-I(M-1); Vb(m)-0; V(m)-O; end end end if (I(m)<-Isat(m)) I(m)-Lambda(m)/L(m); else I(m)-(Lambda(m)-L(m)*Isat(m))/Ls+Isat(m); end end Iterative Calculation of Current and Flux Linkage Profiles % Loop to find corrected current/flux linkage with winding resistance for o-1:5 for m-i:n if Theta(m)<-Thetan Lambda(m)-0; I(m)-0; Vb(m)-0; V(m)-0; Bp(m)=O; Bc(m)-O; dBpdth(m)-O; dBcdth(m)-O; else if (Theta(m)>ThetaOn U Theta(m)<-Thetaaff) Vb(m)-Batt; Lambda(m)-Lambda(m-)+V(m)*(Theta(m)-Theta(m-1))/omega; V(m)-Vb(m) -I (m)*Ract; Bp(m)-Lambda(m)/(N*Ap); Bc (m)-Lambda(m)/(N*Ag); dBpdth(m)-abs((Bp(m)-Bp(m-1))/(Theta(m)-Theta(m-1))); dBcdth(m)-abs((Bc(m)-Bc(m-1))/(Theta(m)-Theta(m-1))); else if (Theta(m)>ThetaOff &k Lambda(m-1)>O) Vb(m)--Batt; Lambda(m)-Lambda(m-i)+V(m)*(Theta(m)-Theta(m-1))/omega; V(m)-Vb(m) -I (m)*Ract; Bp(m) -Lambda(m)/(N*Ap); Bc(m)-Lambda(m)/(N*Ag); dBpdth(m)-abs((Bp(m)-Bp(m-1))/(Theta(m)-Theta(m-1))); dBcdth(m)-abs((Bc(m)-Bc(m-1))/(Theta(m)-Theta(m-1))); else Lambda (m)-Lambda(m-1); I(m)-I (m-1); Vb(m)-0; V(m)-O; Bp(m)-O; Bc(m)-O; dBpdth(m)-O; dBcdth(m)-O; end end end if (I(m)<-Isat(m)) 144 I(m)-Lambda(m)/L(m); else I(m)-(Lambda(m)-L(m)*Isat(m))/Ls+Isat(m); end end end Calculation of Cyclic Work, Power Output, and Losses XEddy Current Power Loss %Eddy Current Power Loss Piston Pep-tlam^2*omega'2*dBpdth. ^2/(12*rhos) *lp*v*hp*4; %Eddy Current Power Loss Stator Pec-tlam^2*omega^2*dBcdth. -2/(12*rhos)*(4*(li+lo)*w*(hc+hv)+4*lw*hw*w); PeCyc-Pep+Pec; Pe-trapz(Theta/omega,PeCyc)*Freq; V.Eddy current power loss total Y.Winding Power Loss PwCyc-I.^2*R; Pv-trapz(Theta/omega,PwCyc)*Freq; XPower Generated net-vork - 1/2*sum(I.*Lambda([2:end,1])-Lambda.*I ([2:end,1D)); Power-abs (net-work*Freq); PowOut-abs (Power) -abs (Pw) -abs (Pe); %Efficiency Eff-abs(PowOut)/abs(Power); 145 Appendix D VRG Component Drawings The two engineering drawings on the following two pages are for the VRG stator and piston laminations. These are the magnetic core components used for testing the VRG system. The components were fabricated by Polaris Laser Laminations. The stator and piston laminations are the essential components to the operation of the VRG system, while the other components fabricate (with the exception of the coil) are support structure components. Dimensions are in inches. 146 6 . IN PR"$HIBR Inslitule 5 IN OR OF THE CONTAINED THIS DRAWING THE PROPERTY OF Mass. of Technology. ANY AS A WHOLE REPRODUCTION PART WITHOUT THE WRITTEN PERMISSION INFORMATION IS SOLE PROPMETARY AND CONFIDENTIAL LO' 0.205 C'4 LO) APPLICATION 1.170 1.380 0.875 10 USED ON A nnealedAI DO NOT SCALE DRAWING FINISH M1 9 29 GaugeSIED MATERIAL TOLERANCING PER: Q.A. COMMENTS: EN A PPR. N INTERPRET GEOMETRIC CHECKED MFG APPR. BEND*2 FRACTONAL0.001 ANGULAR: MACH TWO PLACE DECIMAL 10.01 THREE PLACE DECIMAL 20.001 0.090 2 W G .N NORV Stator Assembly1 SCALE: 2:1 WEIGHT: SZ SHEET 1OF E I Sta tor A sse m b ly C3 LO C) IS 5 PROPRIETARY AND CONHDENTIAL THE INFORMATION CONTAINED IN THIS DRAWING THE SOLE PROPERTYO2 Mass. Institute of Technology. ANY REPRODUCTION IN PART OR AS A WHOLE WITHOUT THE WRITTEN PERMISSION OF Mass. Institute of Technology 15 PROHIBITED. 1.189 4 APPLICATION 1.400 USED ON ----------- UNLESS OTHERWISE SPECIFIED: 31 DRAWN MFG APPR. ENG APPR. CHECKED DIMENSIONS ARE IN INCHES TOLERANCES: FRACTIONAL ANGULAR: MACHt: BEND TWO PLACE DECIMAL t0.01 THREE PLACE DECIMAL 0.001 Q.A. COMMENTS: COEMG.NSE 0.00I INTERPRET GEOMETRIC TOLERANCING PER: MATERIAL FINISH Mn Annealed 3 DO NOT SCALE DRAWING NAME DATE 2 04 0 '9 mT. TITLE REV isto n A sse m b ly A ssembuye Piston Assembly SIZE DWG. NO. SHEET 1 OF 1 Bibliography [1] S. N. Backhaus, Yu Z, and A.J. Jaworski. Design of a low-cost thermoacoustic electricity generator and its experimental verification. 10th Biennial Conference on Engineering Systems Design and Analysis, 2010. [2] Scott Backhaus and Greg Swift. New varieties of thermoacoustic engines. In 9th International Congress on Sound and Vibration. Los Alamos Naitional Laboratory. [3] Scott Backhaus and Greg Swift. A thermoacoustic-stirling heat engine: Detailed study. Acoustical Society of America, pages 3148-3166, 2000. [4] Kees de Blok. Multi-stage traveling wave thermoacoustics in practice. In 19th InternationalCongress on Sound and Vibration. [5] Kees de Blok. On the design of a near atmospheric air operated thermoacoustic engines. 2013. [6] D.E. Cameron and J.H. Lang. The control of high-speed variable-reluctance generators in electric power systems. IEEE Transactions of Industry Applications, 29(6):1106-1109, 1993. [7] J.J. Dineen. Stirling engine displacer gas bearing, 1983. [8] George Dochat. Design study of a 15 kw free-piston stirling engine- linear alternator for dispersed solar electric power systems. Technical report, NASA, 1979. [9] A.E. Fitzgerald, Charles Kingsley, and Stephen Umans. McGraw-Hill, 6th edition, 2003. Electric Machinery. [10] Catherine Gardner and Chris Lawn. Design of a standing-wave thermoacoustic engine. In The 16th InternationalCongress on Sound and Vibration. [11] D.L. Gardner and G.W. Swift. A cascade thermoacoustic engine. Acoustical Society of America, 114(4):1905-1919, 2003. [12] Konstantin Matveev. Thermoacoustic Instabilities in the Rijke Tube: Experiments and Modeling. PhD thesis, 2003. 149 [13] Konstantin Matveev, Andy Wekin, Cecilia Richards, and Najmeddin ShafreiTehrany. On the coupling between standing-wave thermoacoustic engine and piezoelectric transducer, November 11-15,2007 2007. [14] A. Migliori and G.W. Swift. Liquid-sodium thermoacoustic engine. Physics Letters, 53(5):355-357, 1988. Applied [15] Michael Petach, Emanuel Tward, and Scott Backhaus. Design of a high efficiency power source (heps) based on thermoacoustic technology: Final report. Technical report, January 2004 2004. [16] J.W.S Rayleigh. The Theory of Sound, volume 2. Courier Dover Publications, New York, 2nd edition, 1896. [17] Andrea Rossi, Fabio Immovilli, Claudio Bianchini, and Alberto Bellini. Design of linear alternators for thermoacoustic machines. IEEE, pages 2436-2440, 2009. [18] Herbert Roters. Electromagentic Devices. John Wiley and Sons, New York, 1st edition, 1941. [19] Nikolaus Rott. A simple theory of the sondhauss tube. Aeroacoustics, pages 327-338, 1986. Recent Advances in [20] J. Smoker, M. Nouh, 0. Aldraihem, and A. Baz. Energy harvesting from a standing wave thermo-acoustic piezoelectric resonator. Journal of Applied Physics, 111, 2012. [21] Steve Sprague. Lamination steels third edition, a compendium of lamination steel alloys commonly used in electric motors., 2007. [22] Greg Swift. Thermoacoustic engines. Acoustical Society of America, 84(4):11451180, 1988. [23] Greg Swift. Analysis and performance of a large thermoacoustic engine. Acoustical Society of America, pages 1551-1563, 1992. [24] Greg Swift. Thermoacoustics: A Unified Perspective for some Engines and Refrigerators. Acoustical Society of America, 2002. [25] Greg Swift and Scott Backhaus. Gas diodes for thermoacoustic selfcirculating heat exchangers. In A.A. Atchley, V.W. Sparrow, and R.M. Keolian, editors, Innovations in NonlinearAcoustics:17th InternationalSymposium on Nonlinear Acoustics, volume 838. American Institute of Physics Publishing. [26] G.W. Swift and J.J. Wollan. Thermoacoustics for liquefaction of natural gas. Gas TIPS, 8(4):21-26, Fall 2002. [27] O.G. Symko, E. Abdel-Rahman, Y.S. Kwon, M. Emmi, and R. Behunin. Design and development of high-frequency thermoacoustic engines for thermal management in microelectronics. Microelectronics Journal, 35:185-191, 2004. 150 [28] D.A. Torrey and J.H. Lang. Optimal-efficiency excitation of variable-reluctance motor drives. IEE Proceedings, 138(1):1-14, 1991. [29] David Torrey. Optimal-Efficiency Constant-Speed Control of Nonlinear Variable Reluctance Motor Drives. PhD thesis, 1988. [30] Francesco Vallese. Design and Operation of High-Power Variable Reluctance Motor Based Drive System. PhD thesis, 1985. [31] Bill Ward, John Clark, and Greg Swift. Design environment for low-amplitude thermoacoustic energy conversion deltaec, 2012. [32] T. Yazaki, A. Iwata, T. Maekawa, and A. Tominaga. Traveling wave thermoacoustic engine in a looped tube. Physical Review Letters, 81(15):3128-3131, 1998. [33] Zhibin Yu and Artur Jaworski. Impact of acoustic impedance and flow resistance on the power output capacity of the regenerators in travelling-wave thermoacoustic engines. Energy Conversion and Management, 51:250-359, 2009. 151