I‘ :M '9‘, “0"”; d" , 2 " ' . 1;.-.) mtg-«LN. 12:3 ,':5"'5M',.',"" 11%|?“ '55 ""55"qu £ '.'-|'| “:1 5"';..'.""I'1‘"." 2:215" :4 "El . E!“ '.TU"-' :;‘r.:':" . "ELI-H “4'3“ . ' IV ' l N T .'.'.I.’ 5". .‘1 "".5'5' '1." L] ".2“ 95555.3 '1"; ' ."".Ir '.'”"" . '5' 5. .. t ‘5i35-‘N'f'"' 11:31:91.2 "2'32““! ' ' 2"? .112." .1. -_ 2.2-,"1‘: .1352? 2112.1!“ 23121.5” W" q}. ‘2‘ ‘ :96“ ':"'.'.'|"‘ 32.. "'1 25‘ 3| 2:5 . 3". 3 ' 3"59t-'l‘:1!‘: | 1?.3 5112'5'2‘58'J1'ési32'gpqu" Egg. M" 5'1'"""' . .{.... ’ '1': ' d'.|3":‘l‘5<2 ‘8 25.23:.$111511.l9.."'."v"‘..."-,. ' -. "I'555'{:'.' | EM 35.. 5:59:1“H3' "4'5"" '1“ '55:“ '5""'5'.|" ”'5: ant. 5 ' "o‘tLJ'slm L§g‘& l . '5'“? 'i':".'"""' ‘1; H" ”Thur [3212“) 5 ,. 5. “héufir', “ifl't: ' . ' I I. “'5: 11"“ ‘2. . :AI-é '3'”; ' lg?” "r“ ‘ ' .M' 3"...” «m. «1.5L. “ .~ 1.... ‘2' 7333'- :32. "‘ 31.333.5fl"3x°”5"55 3333-. "5 ';""""I '5‘“ I5 “:3 )tlzlq' '- """'1':';' 3 5:3: 3.. 7d. 13%; 3 13 ”1‘55 “11‘ dfii'z‘j" ”in.” .5; 57.1 ' "' ' a ‘5" 'w'fifl 5‘ .' ' " 'O ' 95333:” 545155 :15‘5E1‘5".'!l. "ativ'fi‘kfi‘f- .15; {Evtvq‘ ""111 . J 2‘. . 2.5.1.. 3. . -.. '3'113'..1.5'.i:!“!'.."' .. 5'5'92'3 :7' 2.2 13.1"”: 2|... u... ‘1‘? %'.L»' 22-191 {3‘11” ' "'3‘ .1313}. "1&51‘ | I? 2..! "T.“ 2’1 . .13.“. 25.2335. 5%...53 fl . |'. '33-! .33....'.'; .13.}. {rm}... g3. 51' . n , . 0.. 1.15.4.E‘1Vg‘t" J!" fl- 45‘1"! . L51 15233225121" .5' 2 l “5 :0 5‘5»: 'y-I '; w'i'il '1" 't'55'ifl'x'."'525 “‘fi .22 3 c _ 3, 23 ET? . ' .-.,l ' '5'“ M54 EVEN: ' ' '2 ;. Vic ' " 4‘ I .‘ ‘ .5. 5... '5. : 3.3,.“ .1: £12. W}! 3... 4...... .2: 3315.53“; ' , I. 5 '35:. 12" 0' ”'3 2: .1... 211 l ' '2 -‘ I 'VK... 2 I" J “' 1"5':5:. ""‘ ; "1‘33: 71 '7-.-.' 1.“. 5.2. 2'" :M‘ I ,., ‘ ‘ 1' 1'. 1 7 '1 '5' . :2, ‘ . ’«l . 4's .2 “ '3' ' ""I-JP‘.» .-.- .. n ”I ~11)" ' v 5 L T‘ .3" . I . 2| ~—-v .3- ‘- .n. n ‘ '- (‘p' . _ " o. 97". ",“ - , ..' ”‘1... . . _1 min: 3-23.2' ‘12... 3.3.1.2.. 91W .. n.‘ '9:- I b If 5,1 ~: g, . : ' 921%... ...~ -. .. - D m 5. w -.- ”so. . "M... r?- n~ ~ ” ‘ 5 '0 :2 IlllTTllllllllllllllllllillllllll Kl i 1‘ l 6} 3 1293 01682 2524 This is to certify that the dissertation entitled ANALYSIS AND EXPERIMENTAL STUDY OF RADIATIVE HEAT TRANSFER THROUGH ELECTRORHEOLOGICAL FLUIDS presented by Jeffrey B. Hargrove has been accepted towards fulfillment of the requirements for Ph .D . degree in Mechanical Engineer ing MS U is an Affirmative Action/Equal Opportunity Institution DJ 2771 LIBRARY Michigan State Unlverslty PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MTE DUE DATE DUE DATE DUE M‘- 1m WWpGS-p.“ ANALYSIS AND EXPERIMENTAL STUDY OF RADIATIVE HEAT TRANSFER THROUGH ELECTRORHEOLOGICAL FLUIDS By Jemey B. Hargrove A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering 1997 ABSTRACT ANALYSIS AND EXPERIMENTAL STUDY OF RADIATIVE HEAT TRANSFER THROUGH ELECTRORHEOLOGICAL FLUIDS By J effi‘ey B. Hargrove Radiation heat transfer control utilizing the unique properties of electrorheological (ER) fluids has recently been the subject of considerable interest as an innovative new area of research. While some work has been done to demonstrate the concept and to show the potential for radiation transmittance control, little has been done to Specifically characterize the ftmdamental radiation transport mechanisms involved. This work seeks to identify the dominant modes for attenuation of radiant energy incident upon a typical ER fluid. Models are developed to predict radiation heat transfer through a composite window featuring a central layer of ER fluid while the particles are in a randomly dispersed state. A model for prediction of the enhanced levels of energy transport in the electric field-induced chained particle state was then developed by taking into account particle chain geometries. Additionally, the effect of radiation wavelength was studied for incident light beams ranging fiom 500 nm to 800 nm. Furthermore, the eflect of the angle of incidence of the striking beam onto the composite window was studied, and a model is proposed to predict the change in transmittance associated with the angle of incidence. The levels of transmittance predicted by these models were compared to data obtained by experimental measurement, and excellent agreement was shown. The result of this work is a set of models for radiative heat transfer in ER fluid-based composite windows, both in the particle-dispersed and particle-chained states for cases when the plane of the window is held normal to the path of the incident light beam and when it is off normal by some angle, for wavelengths of incident light ranging from 500 nm to 800 nm. Clearly, an understanding of the physical mechanisms involved will provide insight into understanding heat transfer augmentation in the other primary modes. The models developed can be used to develop sensors used for optimal control strategies to manage the response of ER fluids for reliable use in commercial and industrial applications. Furthermore, with reliable control of the heat transfer properties of ER fluids will come the opportunity to develop thermally “smart” materials that will provide solutions to advanced heat transfer problems encountered in next generation technologies and applications. This work is dedicated to my wife Laurie whose patience exceeds words of admiration iv ACKNOWLEDGMENTS I would like to express my gratitude for the faithful guidance of my advisor and friend, Dr. John R. Lloyd. Throughout my entire graduate program he has served to guide not only my academic development, but also my professional development from a philosophical standpoint that can only be thought of in terms of legacies and excellence in commitment to developing students. I hope that I may someday reflect his distinct qualifications to those that I mentor. The assistance of my doctoral committee members, Dr. Clark Radcliffe, Dr. James Beck, Dr. Craig Somerton, Dr. Henry Kowalski and Dr. Roger Callantone is also gratefully acknowledged. The collective insight that each of these individuals gave me played a critical role in the shaping of this work. A fond note of thanks is also given to a number of my fellow graduate students for their support and assistance in this endeavor. Thanks to Gloria Elliot, Ruth Andersland and Stefan Tabatabi, just to name a few. For all of the members of my families, especially my parents, I offer my most sincere words of gratitude. Great appreciation is given to my colleagues at GMI Engineering & Management Institute in Flint, Michigan. I’m especially grateful to the persistent encouragement of my long-time mentor and friend, Professor Gary Hammond. Finally, a fond note of thanks is due to all of my dear friends for their endless support and encouragement. For all who listened and understood, although far removed from the intricacies of the problem at hand, you are beautiful people. vi TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES NOMENCLATURE CHAPTER 1 - INTRODUCTION 1.1 BR Fluid Fundamentals 1.2 Objectives and Scope of This Work 1.3 Overview of ER Fluid Technology CHAPTER 2 - EXPERIMENTAL APPARATUS AND METHODS 2.1 Experimental Apparatus 2.2 Experimental Methods CHAPTER 3 - RADIATIVE TRANSFER THROUGH RANDOMLY ORIENTED PARTICLES 3.1 Extinction of Radiant Energy 3.2 Experimental Results & Discussion CHAPTER 4 - RADIATIVE TRANSFER THROUGH PARTICLES IN THE FIELD-INDUCED CHAINED STATE 4.1 Additional Experimental Apparatus 4.2 Analysis of Particle Chain Geometry 4.3 Experimental Results and Discussion CHAPTER 5 - ANGULAR TRANSMITTANCE ANALYSIS 5.1 Additional Experimental Apparatus 5.2 Analysis of Particle Chain Geometry at Oblique Angles vii ix xii Ox-hsr—v— l4 l4 19 25 26 32 41 41 42 46 58 59 61 5.3 Experimental Results and Discussion CHAPTER 6 - OPTICAL SENSORS FOR STATE CLASSIFICATION CHAPTER 7 - CONCLUSIONS AND RECOMMENDATIONS FOR FURTHER STUDY APPENDIX A - AN OVERVIEW OF RECENT DEVELOPMENTS IN ER FLUID TECHNOLOGY APPENDD( B - EXPERIMENTAL DATA FROM AN GULAR TRANSMITTANCE MEASUREMENTS APPENDD( C - EXPERIMENTAL UNCERTAINTY ANALYSIS C.l Sources of Error C.2 Error in the Dispersed Particles Model C.3 Error in the Chained Particle Model C.4 Error in the Chained Particle Model With Field Strength as a Parameter C.5 Error in the Angular Transmittance Model BIBLIOGRAPHY viii 71 78 82 85 93 96 96 100 101 103 105 109 LIST OF TABLES Table 3-1 Experimental vs. Calculated Extinction Coefficients (Dispersed Particles, 2a = 11pm, L = 1.9 mm) Table 3-2 Experimental vs. Calculated Extinction Coefficients (Dispersed Particles, fv = 0.005, L = 1.9 mm) Table 4-1 Experimental vs. Calculated Transnrittance Levels (V; = I 6 0 Vrms/mm, L = 1.9mm, 20 = 11 pm) ix 37 38 54 LIST OF FIGURES Figure 1-1 Conceptual Schematic of ER Fluid Particle Dispersion Under No Field 3 and Chain Structure with a Sufliciently Large Field Applied Figure 2-1 Schematic of Experimental Apparatus 14 Figure 2-2 Photograph and Schematic of the ER Fluid Window Showing the 16 Components Held in Place by the Holding Fixture Figure 2-3 Photograph of the Experimental Apparatus 17 Figure 2-4 Transmittance Through a 50-50 Beam Splitter and a 10 Percent 18 Transmitting Neutral Density Filter Figure 2-5a Zeolite Particles, Average Diameter 11 Microns 20 Figure 2-5b Zeolite Particles, Average Diameter 25 Microns 21 Figure 2-5c Zeolite Particles, Average Diameter 40 Microns 21 Figure 3-1 Cross Section of Zeolite Particle Showing Radiation Absorbing Cavities 29 Figure 3-2 Normal Transmittance Through ER Fluids of Different Volume Fractions 33 with Fully Dispersed Particles (L = 1.9 mm, 2a = 11 ,um) Figure 3-3 Transmittance as a Function of Sensor Position Transverse to the Axial 34 Direction of the Incident Beam Figure 3-4 Transmittance in the Dispersed State for Varying Volume Fractions as a 36 Function of Pathlength (Za = I 1 pm) Figure 3-5 Transmittance in the Dispersed State for Varying Particle Sizes as a 38 Function of Pathlength (fv = 0.005) Figure 3-6 Comparison of Experimental Transmittance to the Fundamental Beer’s 39 Law (2a = 11pm) Figure 4-1 Microscope Electrode Apparatus 42 Figure 4-2 Uniformly Spaced Single-Particle Chains for the Ideal Case of F ield- 43 Induced Chaining Figure 4-3a Stacking Effect of Particles at the Electrode 47 Figure 4-3b Microscopic Video Still Showing Stacking Effect 48 Figure 4-4a Early Stages of Chain Bending 49 Figure 4-4b Early Stages of Chain Bending 49 Figure 4—4c Initial Contact Between Chains 50 Figure 4-4d Initial Contact Between Chains 50 Figure 4-4e Single Column of Multiple Chains 51 Figure 4-4f Single Column of Multiple Chains 51 Figure 4-5 Transnrittance in the Chained Particle State as a Function of Volume 54 Fraction (V;= 160 Vrms/mm, L = 1.9 mm, 2a =11,um) X Figure 4-6 Effect of Field Strength on Transmittance (fv = 0.0075, L = 1.9 mm, 2a = 11 .10") Figure 4-7 Efl‘ect of Field Strength on Transmittance - Linearized Model (fl, = 0.0075, L = 1.9 mm, 2a = 11 ,um) Figure 5-1 ER Fluid Window with Normal Energy Incidence, Viewed from the Top Figure 5-2 Rotation of ER Fluid Window About a Vertical Axis Figure 5-3 Experimental Apparatus for Angular Transmittance Measurements Figure 5-4 Particle Chains Parallel to the Path of Incident Energy Figure 5-5 Particle Chain Profile in a Rotated ER Fluid Window Figure 5-6 Blocking Area of Particles That are Exposed to Incident Energy Figure 5-7 Conceptualization of a Rotated Particle Chain in 3d Figure 5—8 Effective Blocking Area Figure 5-9 Overlapped Columns Extending Across the Entire Area of Irradiation Figure 5-10 Angular Transmittance: Experimental Data Versus Model Predictions (L = 1.9 mm, Za = 11 gm, Vf= I60 Vrms/mm) Figure 5-11 Transmission Intensity of Energy Through Glass as a Function of Angle Figure 5-12 Transmission Intensity of Energy Through Glass as a Function of Angle Figure 5-13 Sampled Transmitted Energy Data Points Against a Model of the Waveform of Incident Energy xi 55 56 58 60 61 62 63 65 66 68 69 73 74 75 76 AB Cabs C ext NOMENCLATURE mean particle radius particle volume fraction absorptive index index of refraction pathlength variable direction unit vector particle size parameter total blocking area of all particle chains area of blocking particle-chain face absorption cross section extinction cross section volume fraction dependent total number of columns scattering cross section number of chains per volume of ER fluid incident energy incident energy absorbed incident energy transmitted through an empty window incident energy scattered xii Qext Vf 0). Ta Tc 9: calculated average extinction coeflicient the extinction pathlength of ER fluid effective pathlength number of particles per unit volume total number of particles per volume of ER fluid extinction efficiency volume of ER fluid RMS voltage field strength per millimeter volume of an individual particle extinction coeficient absorption coefficient incident radiation wavelength scattering coeflicient transmittance in the dispersed state transmittance in the chained state azimuthal angle angle of transmission through glass xiii CHAPTER 1 - INTRODUCTION 1.1 ER Fluid Fundamentals In 1947, Willis M. Winslow introduced the concept of changing the rheological properties of certain suspensions of particles by subjecting them to an electric field of sufficient strength (Winslow, 1947). These substances have since been termed “electrorheological (ER) fluids”, and since the time of the very earliest descriptions of the phenomena associated with them, the scientific community has actively sought to find ways to capitalize on them. Over the years, countless numbers of researchers have studied the phenomena of ER fluids. From their foundational work, many theories have emerged and have been refined that have helped to shape our understanding of the fundamental physical mechanisms behind these responses. Along the way numerous devices and commercial applications have been proposed. Yet, for all of the potential held by devices that would employ ER fluids, engineering and design limitations continue to persist that have kept applications elusive for the most part. There is a specific void in understanding the response of ER fluids to an applied electric field and in achieving enough repeatability and reliability in that response to confidently utilize them in applications. This work will seek to further that 2 understanding, and to provide for a unique way to provide a measure of the geometric micro-state of the ER fluid’s particles by quantifying the level radiative of energy transmittance through it. As this is achieved, the development of optical sensors will enable development of optimized feedback control systems, as a key to managing the response of an ER fluid, can proceed. Therefore, this work will have far reaching ramifications on the future of ER fluid-based applications, and is critical to progress in a number of ER fluid related fields. Electrorheological fluids are suspensions of highly polarizable, micron sized particles suspended in a suitable carrier fluid. Typically, the particulate medium is hydrophilic, whereas the carrier fluid is hydrophobic. Ideally, the Specific gravity of the particulate would match that of the carrier fluid, so that maintenance of long term dispersion is ensured. When placed in an electric field of sufficient strength, the otherwise randomly dispersed particles form particle chains that are aligned with the field, thus giving rise to a complex “fibrous”, or chained, structure as illustrated in Figure 1-1. Depending upon the strength of the field, this chaining structure will advance toward a body centered cubic orientation, which represents the lowest possible energy state (Tao and Sun, 1991). Additionally, upon removal of the electric field there is a tendency for thermally induced motion to overcome the weak colloidal forces holding the chains together, thus allowing the particles resume their random distribution. Hence, the process of electrically-induced chaining is reversible without need of an external stimulus, but as we will find out, not necessarily fast. Figure l-1 Conceptual Schematic of ER Fluid Particle Dispersion Under No Field and Chain Structure with a Sufficiently Large Field Applied The immediately obvious areas of application are those that capitalize on the mechanical properties of ER fluids, such as viscosity. When the particles are in the field- induced chained state, the apparent viscosity of the ER fluid is dramatically increased, suggesting use in hydraulic valves, mechanical transmissions, and active suspension systems (Morishita and Ura, 1993). But more recent work has suggested an entirely new and revolutionary area of application for ER fluids in the realm of augmenting and controlling heat transfer in thermal processes (Shul’man, 1982). Zhang and Lloyd (1992, 1994) have shown that radiative transmittance in ER fluids is significantly increased when the particles are in the field-induced chained state. Their work, along with supporting work by Andersland (1996), has also shown that transmittance levels can be managed by controlling the 4 electric field strength. Both studies have demonstrated these concepts with strongly supportive experimental data, but have only suggested a theoretical basis for understanding the physical processes affecting radiative transfer. 1.2 Objectives and Scope of This Work In the introductory statement, it was observed that commercially successful ER fluid-based applications have been elusive. This may be due, in part, to the extremely interdisciplinary nature of ER fluid technology. To fully understand the nature and flow properties of ER fluids will require an understanding of the physics of fluids, particulate systems and solids, colloid and surface chemistry, electrochemistry, mechanical engineering, electrical engineering, mathematics, polymer and material science, just to name a few. Only recently have trends in engineering research been such as to bring together individuals of diverse expertise with a singular outcome objective. Naturally, increased fundamental knowledge of the physical mechanisms of ER fluid response to applied fields is a necessary forerunner to applications. In order to determine the feasibility of any device utilizing ER fluids, many key parameters must be known for a wide variety of operating conditions and parameters. This includes, but is certainly not limited to, thermal conditions, aging considerations and moisture content of the particles of the ER fluid, and a wide array of applied voltage waveforms and parameters that the ER fluids may be subjected to. Furthermore, this entire process is complicated by the vast number of physical parameters associated with the ER fluid itself. 5 Varying particle properties and carrier fluid properties must be accounted for and then related to all other system parameters before optimal responses may even conceivably be achieved. The overall objective of this work is to further advance ER fluid related technology toward applications. Specifically, this work will entail three interrelated studies that contribute to providing an understanding of radiative heat transfer through a composite window featuring a layer of ER fluid made of silicone oil and Zeolite particles. The first study focuses on the modeling of radiative heat transfer through the condition where the ER fluid’s particulate medium is in a fully dispersed state. The second study then considers radiative heat transfer for the condition where the particles are in an electric field-induced chained state. In each of these first two studies, the beam of incident radiation is applied to the window normal to the plane of the window. In the third study the window is rotated about a vertical axis passing through the mass center of the ER fluid so that off-normal transmittances may be quantified. In all of the studies, a sequence of experiments was conducted in which transmittance measurements were taken over a range of volume fractions, particle sizes, extinction pathlengths and electric field strengths. The experimental data collected were used to make comparisons to levels of transmittance predicted by the newly developed models, and hence provide a measure of the validity of the models. 1.3 Overview of ER Fluid Technology To review the history of progress and advancement of ER fluid technology, one could easily examine the past half-century of research and published findings and gain a thorough understanding of the phenomenon. There has been a tremendous amount of research on the subject since the findings of Winslow (1947), and acknowledgment of the contributions of many investigators should be made. However, to exhaustively survey the past work would not necessarily provide any insight or contribution to this thesis. Instead, this section will focus on providing an overview of recent work that has a more direct significance to the scope of this investigation. Altemately, to provide the reader with the opportunity to review the most recent developments in ER fluids research, Appendix A is provided. A more complete survey of the past few year’s research in all areas of ER fluids technology is found there. Within these works one can also gain a general overview of the full field of ER fluid technology, and gain reference to the earlier pioneering works of those who have brought this field of technology to where it is today. An overview of recent progress in the field of ER fluid research should begin with acknowledgment of the great deal of effort that continues to be applied toward understanding the basic mechanism of the phenomenon of ER fluids. The advancement of ER fluid technology of this investigation, as any other, critically depends on good firndamental knowledge of the physics behind ER fluids as its basis. 7 One such study to understand the fundamentals of ER fluids (Kita et a1, 1993) cites that although numerous theoretical and experimental studies have been conducted at many laboratories and institutions, there is still a controversy regarding the fundamental mechanism involved in the ER efl‘ect. In this investigation, the effects of electric fields, shear rates, and particle concentration on shear stress acting in the ER fluid were studied. This led to classification of fluid flow between electrodes as either (1) an almost-solid- state-layer, (2) a two-phase layer, or (3) a pure-liquid layer. In a study regarding the role of water capillary forces in ER fluids (See et al, 1993), the authors point out that many researchers have noted that the viscosity of ER fluids under electric fields increases through a maximum as water content of the particles is increased. The authors explain this by suggesting that the particles are held together by the capillary force of water trapped in their contact area. A model is proposed that equates cohesive force between particles to numerous microscopic water bridges created by electric field. The number of these water bridges depends on field strength and amount of adhered water. Experimental confirmation of the model, along with estimation of time scales for formation and relaxation of bridges is presented, along with comparable experimental results. The effect of electrical field characteristics on static shear stresses in ER fluids has also been the subject of a recent study (Ginder and Ceccio, 1995). Here, the authors investigate the effect of time-varying electric fields on the stresses in a variety of ER fluids in steady shear. By studying the dependence of the time-averaged shear stress on the frequency of the applied field, two classes of ER fluids were realized: (1) those that 8 exhibit considerable ER shear stresses with DC and low frequency AC fields, and (2) ER fluids that perform poorly in DC fields but show considerable electrorheological activity in AC fields. The authors fruther point out that these behaviors are consistent with the existence of two separate means of producing ER activity, that is, mismatches of conductivities and dielectric constants, respectively, between the suspended particles and the suspending fluids. In this study, it was further noted that the characteristic times for the growth and decay of induced stresses varied only weakly with strength of the applied field. It was also shown that response times were shortened with increased shear rates. In yet another study on the effects of AC and DC electric fields, the dynamic shear response of ER fluids consisting of rod-like particles was investigated (Kanu and Shaw, 1994). Here, the authors propose that particle geometries should be considered to enhance the particle-particle interactions that are typically accepted as being solely responsible for the rheological properties of ER fluids. Experimental evidence shows that the dynamic modulus increased with increased particle aspect ratio. The influence of the orientation of the electric field on shear flow of ER fluids has also been studied (Ceccio and Wineman, 1994). This work considered the possibility of electric fields with components both normal and parallel to the direction of fluid velocity. The analysis done included the effects of the electric field on (1) the material parameters and (2) the contribution to stress components because of the interaction of the electric field vector and the shearing. There has also been a recent study on the eflect of prior heating of Zeolite/silicone oil ER fluids on their subsequent electrorheology (Conrad et al, 1994). BR fluids were 9 heated to 80 degrees Celsius and then cooled to 25 degrees Celsius. Results showed that heating generally produced a decrease in permittivity and current density. Furthermore, it was shown that an increase in shear resistance after being heated to up to 80 degrees Celsius could be expected, but at higher temperatures a decrease in shear resistance occurs. It was concluded that both a decrease in water content and a rearrangement of water remaining in the Zeolite cages was the cause of the mechanical changes experienced. Another broadly studied area related to the basic understanding of the ER fluids phenomenon was that of sensing the structure of the field-induced particle chains. This is of particular interest to this investigation. A study of permittivity of ER fluids under steady and oscillatory shear (Adolf et a1, 1995) was used to show a relationship between viscosity of the ER fluid suspensions and permittivity for a variety of volume fiactions and applied fields. This work showed that the highest permittivities coincide with the lowest levels of strain under high fields and low strain amplitudes, while coinciding with the minimum strain rate under low fields and high strain amplitudes. In the time dependent dielectric response of quiescent ER fluids (Adolf and Garino, 1995), which was a follow-up study to the one above, the evolution of particle structure was monitored by following the increase in perrrrittivity after an electric field was applied. Measured permittivity agreed well with predictions for dense columnar aggregates for low particle volume fractions and high fields. However, the authors point out that for higher concentrations of particles or lower fields the measured perrrrittivities were lower than expected because final particle chain structures were less compact. In 10 another work with similar objectives, normal stress measurements have been used to indicate the induced chain network structure (Kimura et a1, 1994). Ginder (1993) has used diffuse optical probes to measure transmittance through a commercial ER fluid in an effort to observe particle motion and structural formation. This work has successfully provided a measure of the time scale for field-induced structure formation, by measuring the temporal variation in diffuse transmittance, that is consistent with the induced-dipole attraction between particles. The dependency of response times on field strength is studied in this work as well. As indicated in Appendix A, there continues to be tremendous effort and resulting strides being made in applied research that seek to capitalize on the unique mechanical properties of ER fluids. However, a noticeable level of research focused on taking advantage of the equally unique thermal properties of ER fluids is starting to emerge. Beginning with a pioneering study by Shul’man (1982), and subsequent works, an overview of enhancements in therrnophysical properties by applying electric fields to ER fluids has been attained. For example, it has been conclusively shown that increases in heat conduction by as much as 55 percent are achievable (Zhang and Lloyd, 1993) while the Specific heat of the ER fluid remains unchanged. Further studies have quantified the characteristics of ER fluids flowing in an axial direction between concentric cylinders (Tsukiji and Furuse, 1994). Here, pressure drop and flow rate were measured under application of electric field. In work of a similar nature to that presented in this thesis, the optical effects of ER fluids were studied by Hunter et a1 (1993). The authors reported a transmission drop 1 1 when fields were applied perpendicular to the light path. Furthermore, the authors report no drop in transmission for fields parallel to light path, which would otherwise contradict the reportings of other investigators who have shown increases in transmission for such fields. The authors also report that they were unable to resolve the dependence on the electric field. Several studies have been carried out that treat light transmission through ER fluids as a scattering phenomenon. Martin et a1 (1993), Halsey (1993) and Martin and Odinek (1995) have performed a comprehensive study of light scattering through the evolving ER fluid particle structure during transition upon application of an electric field. Some results of these studies include prediction of increasing column width with time, where the time constant decreases with increased field strength. Another result was the development of a theoretical model describing the dynamics of field-induced structures in oscillatory shear based on light scattering observations and by considering the efi‘ects of hydrodynamic and electrostatic forces on a fragmenting / aggregating particle chain. Finally, in a study that used a technique that proved critical to the work presented in this thesis, Fisher et al (1993), used optical microscopy and statistical analysis to quantitatively determine the structure of glass bead ER fluids as a fimction of electric field strength and concentration of beads. The authors note that the initial structure was of single particle (bead) chains, and that increased fields caused only slight changes in the structure. These observations parallel those made in this work. In the previous study, the authors note that no clear effect of bead size on chain structure was observed. 12 The work of this thesis is an extension of recent work that has suggested an entirely new and revolutionary area of application for ER fluids in the realm of augmenting and controlling heat transfer in radiative processes. Zhang and Lloyd (1992, 1994) have shown that radiative transmittance in ER fluids is significantly increased when the particles are in the field-induced chained state. In their work, a monochromatic beam of radiation energy was transmitted through a composite window cell containing a layer of ER fluid. Radiation heat transfer was studied under various physical conditions that included particle volume fraction, electric field strength and extinction pathlength. In addition to measuring significant increases in transmitted energy upon application of an electric field, the authors were able to identify the primary controlling parameters in the response of the ER fluid. More recently, supporting work (Andersland, 1996 and Radcliffe et al, 1996) has shown that transmittance levels can be managed by controlling the electric field strength through the uses of sensory feedback. By employing feedback control, response times of particle alignment upon application of an electric field were shown to be 35 times faster than those achieved by direct application of the electric field without any form of feedback. In addition, it was shown that desired transmittance levels could be achieved using feedback control with dramatically accurate results. Both studies have demonstrated these concepts with strongly supportive experimental data, but have only suggested a theoretical basis for understanding the physical processes affecting radiative transfer. As stated earlier in the scope of this work, a primary objective of this work is to 13 establish the underlying mechanisms behind radiative heat transfer, and to provide models to describe such processes. To do so, this work will focus specifically on identifying the fundamental modes for attenuation of radiant energy through a typical ER fluid. As this is completed, the focus will shift to (1) development of a model to describe transmittance through an ER fluid-based composite window in which the particles are in a randomly dispersed state, (2) development of a similar transmittance model for the chained-particle state, and (3) development of a transmittance model for the case where the path of incident radiation is not parallel, or at an oblique angle, to the direction of particle chains. CHAPTER 2 - EXPERINIENTAL APPARATUS AND METHODS 2.1 Experimental Apparatus A schematic diagram of the apparatus used to take all experimental data is seen in Figure 2-1. Note that the apparatus as shown shows the plane of the window being held normal to the path of the incident beam of radiation energy. A variation to this was done for the third study, where the window was rotated about an axis through the mass center of the ER fluid (perpendicular to the page) to provide off-normal transmission of energy. ER Fl 'd Light Beam Wind; SPEC T ROME T ER 3 LIGHT SOURCE PC - BASED J Beam Shown Passing Through ER Fluid Window at Normal Incidence - HIGH VOL TA GE GENERA TOR FUNCTION I GENERA TOR Figure 2-1 Schematic of Experimental Apparatus 14 15 The window consisted of two pieces of glass separated by a gasket, as illustrated in Figure 2-2. The center of the gasket was cut out to form a chamber into which the ER fluid could be injected. The glass used was 1 mm thick soda lime glass coated with a 900 angstrom thick layer of indium-tin oxide (ITO). The gasket material was sulfur-cured styrene butadiene, which served two distinct purposes: (1) to create a seal around the glass so that the ER fluid did not leak out, and (2) to act as an electrical insulator capable of withstanding the typically high field strengths associated with ER fluid use. All of this was mounted in a steel holding fixture that was designed specifically to hold the window at a desired orientation to the incident energy beam. Because of the nature of the experimentation, where the ER fluid window was repeatedly removed from the holding fixture and replaced with another experimental window, the holding fixture was designed for extreme precision and repeatability of window location. This was accomplished by implementing rigid stops at all points where variation of window location could occur. Furthermore, the holding fixture was electrically insulated from the window by a layer of silicon rubber placed between the glass and the steel. For the transmittance experiments, it was desired to measure transmittance through a window in which the beam of energy’s incidence angle was normal to the plane of the window. The holding fixture was also used for mounting the optics to present the incident radiation, as is described in greater detail below, and was therefore designed so that the incident radiation beam was always oriented normal to the composite window. 16 Steel " Insulator Insulator l” l l“ 7.2., ,; l 1 )l ,, «Iii 'i .. r Actual Photograph 1 .‘l '7' H" ‘ ,_ '11:} is“ G ass ofthe Holder ,aj-li i ii...) liiri’Gasket Assembly [11‘ ‘ . W allfTLGlass // "\j l i ‘ ‘l ‘ I” g 1 Steel Figure 2-2 Photograph and Schematic of the ER Fluid Window Showing the Components Held in Place by the Holding Fixture The electrically conducting ITO coating served as an electrode to distribute the electric field uniformly across the window glass and through the ER fluid. The coating, which had a thickness of about 900 angstroms, was thin enough so that the amount of beam energy lost to absorption by the coating was minimal. Normal transmittance through glass alone (no coating, two pieces) was measured to be 96.5 percent. The addition of the coating on the two pieces reduced the transmittance to 93.5 percent, meaning that on an individual piece of glass the coating reduces transmittance by 1.5 percent. Regardless, the overall reduction in transmittance was mathematically eliminated from measurements involving the ER fluid, as described at the end of this section and in section 2.2. Figure 2-3 shows a photographic layout of the entire experimental apparatus. Figure 2-3 Photograph of Experimental Apparatus The major components include a PC-based spectrometer. The PC1000 manufactured by Ocean Optics was used for taking transmittance data. This was connected to an 80486 nricroprocessor based PC. SpectraScopeGD version 2.3 software from Ocean Optics was used to process the spectral data, including radiant intensity, from the spectrometer. A trmgsten-halogen lamp served as the light source, providing useful light in the wavelength range of 500 nm to 800 nm. This incident radiation was transnritted to the ER fluid-based composite window via a 400m fiber optic cable with a 3 millimeter diameter collimating lens attached to its end. Radiation transmitted through the window was collected by an identical collimating lens, which was aligned to the path of the incident energy beam, and focused on another 400nm fiber optic cable that was connected to the input of the spectrometer. Both collimating lenses, the temrinal ends of 18 the fiber optic cables, and the composite window were affixed to rigid steel holders that were designed so that their location and orientation could be maintained for repeatability. Before shipment fiom its manufacturer, the PC 1 000 spectrometer was calibrated to a standard low pressure mercury-argon light. Calibration coefficients were provided for a regression fit that the software uses to determine the true spectral nature of the incident light. After installation in the PC, and before using the instrument for measurements, calibration was verified. This was done by observing the transmittance of light passing through a 50-50 beam splitter and a 90 percent attenuating (10 percent transmitting) neutral density filter. It should be noted at this point that these levels of attenuation (50% and 90%) should only be treated as average values. Figure 2-4 shows the actual transmittance of the tungsten-halogen light source over the observable wavelength range. 1 0'9 1” [Wavelength (A) - nmJ 0.8 + 0.7 ~- 50-50 Beam splitter § 0.3 .. E 05 1M """"""""""""""""""" W' 0.4 «» X *- 0.3 T Expmm' ental Values 02 >A Q1 W - ............................ q 0 agezgaraaagsagtgaas §a§3$§§§§§ §§~33§E§ mutton Vthvelength Figure 2-4 Transmittance Through a 50-50 Beam Splitter and a 10 Percent Transmitting Neutral Density Filter. 19 The dashed horizontal lines indicate 50 and 10 percent transmission. The upward drift observed is due to the slight decrease in absorptance in the filter expected with larger wavelengths (Modest, 1993). The average value of transmittance is 8.8 percent (12% error) through the neutral density filter and 48.6 percent (2.8% error) through the beam splitter. As a further check, other devices used for light transmittance measurement, as described by (Zhang and Lloyd, 1992 & 1994) and (Andersland, 1995) were used to independently verify transmittance levels as determined from the spectrometer data. Prior to all transmittance experiments with ER fluids, transmitted intensity levels through the glass windows filled with carrier fluid only were measured and recorded by the Spectrometer. This was done to isolate the effect of the carrier fluid and window coating on transmittance, and focus attention on the particles as the cause of extinction. 2.2 Experimental Methods The electrorheological fluids used in this study consisted of particles of anhydrous crystalline Zeolite. The average particle sizes (diameters) used were about 11, 25 and 40 microns respectively, as shown in Figures 2-5a, 2-5b and 2-5c. As can be seen best in Figure 2-5b, which is an electron-micrograph of the 25 micron diameter Zeolite, the particles are made up of conglomerations of crystals, with individual crystal sizes typically 1-3 microns. To insure minimal water adsorption, and repeatability of experiments, the particles were baked at 300°F for a period of time sufficient to vaporize any attached water. Monitoring of particle weight was used while the baking process was 20 done to insure that no significant water remained, as indicated by a stabilization in weight. These particles were mixed with phenylmentlryl polysiloxane silicone oil in varying volume fractions to create the ER fluid. To insure good random dispersion of the particulate, the suspension was stirred prior to each experiment for 10 minutes by a magnetic stirrer, and was then allowed to set for 10 minutes to allow air bubbles created by the stirring process to rise to the surface and dissipate. ISKU 82.090 Figure 2-5a Zeolite Particles, Average Diameter 11 Microns 188“» 682205 Figure 2-50 Zeolite Particles, Average Diameter 40 Microns 22 A syringe was then used to inject ER fluid into the cavity. Once the cavity was filled with ER fluid, transmitted intensity was measured while the particles were in their randomly dispersed state. Transnrittance levels were calculated from the data by simply dividing transmitted energy through the ER fluid by the transrrritted energy through the window filled only with silicone oil at the same wavelengths. All transmittance levels presented were taken from data at 684 nm, which was representative of the location of average transmittance levels over the entire spectrum. In addition to the particle sizes already mentioned, experiments were canied out for a range of particle volume fractions for the fluids that varied from 0.0025 (0.25 percent) to 0.0150 (1.5 percent). Pathlengths were also varied between values of 0.7 mm and 2.2 mm, giving a comprehensive set of data for a wide variety of transnrittances as a fimction of pathlength, particle size and volume fraction for both the dispersed state and the field-induced chained state. For the field-induced chained state, field strengths were varied from less than 100 RMS (root-mean-square) volts per millimeter (V rmsImm) to about 500 Vans/mm. This AC signal had a fi'equency of 60 Hz. The electric fields were provided by amplifying the signal from a Goldstar model FG-2002C function generator by a Trek model 609C-6 high voltage amplifier that had a fixed gain of 1000. Signal amplitudes and frequencies were verified by a Hewlett Packard model 1222A oscilloscope prior to each experiment. When a voltage field was applied to the ER fluid window, a period of time of 30 minutes was allowed to transpire before transmittance data was taken. This was to allow ample time 23 for a steady state condition of particle chaining to be achieved. It should be pointed out at this time that heating of the ER fluid, both from the incident energy beam and fiom resistive heating upon application of the electric field, was negligible. A thermocouple applied to the ER fluid window’s surface took continuous measurements of temperature over the span of time that the experiments were in progress. At no time was an increase in temperature greater than 0.1 degree Celsius observed. For the third study, the window was rotated about a vertical axis through the mass center of the ER fluid held in the window. This is described in greater detail in chapter five. For this study, the ER fluid window assembly and holding fixture were mounted to a Zeiss model 6334 rotary indexing table. Again, to insure precision and repeatability, the holding fixture was mounted to the indexing table with rigidly mounted locating pins. In this experiment, transmittance through the window that was rotated as much as - 12.0 to +120 degrees away fiom normal was measured. In all experiments a set of standard or control parameters were used. The standards used were a volume fraction of 0.005, particle diameter of 11 microns, pathlength of 1.9 millimeters, field strength of 160 Vrms / millimeter. Furthermore, with the exception of the angular transmittance study, the energy beam was held at a 90 i 0.5 degree angle of incidence (normal) to the plane of the ER fluid window. This level of repeatability of the angle of incidence was determined by measuring the variation in location of the reflection of a laser beam aimed at the window. The overall accuracy of the measured transnrittances in the described experiments is estimated and comprehensively presented in Appendix C. This estimation of 24 uncertainty includes various error sources introduced in the experiments such as variation in volume fractions, measured pathlengths, particle diameters, field strength and initial transmission of light energy on an empty window. CHAPTER 3 - RADIATIVE TRANSFER THROUGH RANDOMLY ORIENTED PARTICLES As already noted, radiation heat transfer control utilizing the unique properties of electrorheological (ER) fluids has recently been a subject of interest as an innovative new area of research. While much work has been done to demonstrate the concept and show the potential for radiation transmittance control, little has been done however to specifically identify the fundamental radiation transport mechanism involved. This chapter seeks to determine the dominant modes for attenuation of radiant energy incident upon an ER fluid made of Zeolite particles. Specifically, the focus of this chapter will be on the case where the particles are in a fully-dispersed, or randomly oriented, state. Such would be the case in the ER fluid prior to application of the electric field. Once the dominant modes of attenuation are established, appropriate models will be developed to predict radiation heat transfer through a composite window featuring a layer of ER fluid. Later on in the chapter, a comparison will be made between the levels of extinction predicted by these models and data obtained by experimental measurement. It is known that when the particles of an ER fluid are in the fully dispersed state under a zero-voltage condition, each particle acts to attenuate incident radiation. Upon field-induced chaining (see Figure 1-1), small trmnel-like regions form that are void of 25 26 particles. This allows a greater amount of radiant energy to pass unimpeded, hence increasing transmittance (Tabatabai, 1993). In attempting to quantify attenuation, it has been frequently stated that each particle may be treated as a point scatterer. Assuming that the multiple scattering regime has not been entered, the classic Mie solution for light scattering (van de Hulst, 1957; Kerker, 1969; Brewster and Tien, 1982; Buckius, 1986) could be employed to define radiative properties for transfer through an ER fluid. Yet, the distinctly different mechanisms of scattering and absorption have often been lumped under the singular term extinction, and little has been done to differentiate between the true cause of attenuation for ER fluids. The obvious benefit of doing such a study is understood when one considers the inherent mathematical difficulties involved with determining scattering parameters. Scattering is the result of mutual enhancement and cancellation of reflected wavelets of energy (Bohren and Huffman, 1983). Modeling the attenuation of incident energy transport would be greatly simplified if the particles involved could be treated as point absorbers that Simply remove energy from the beam with no further inter-particle interaction. 3.1 Extinction of Radiant Energy Radiant thermal energy transport in an absorbing, emitting, and/or scattering medium is described by a radiative energy balance. It may be stated that “the change in 27 radiation intensity along a path defined by a vector «3 through a medium is due to a positive contribution fi'om emission and fiom radiation scattered into a“, minus contributions fiom absorption and scattering away from the a‘ direction. ” (Modest, 1993). This energy balance leads to various forms of the equation of radiative heat transfer. Since operating temperatures involved with ER fluids are typically low, the contribution due to emission may be neglected. Absolute absorption is proportional to the magnitude of incident energy and the distance the beam travels along a pathlength 0 _<_ s S L through the medium. The change in intensity of an incident beam in the distance ds is thus written as d] =—K,Ids (3.1) abs where the negative Sign denotes a decrease in intensity, and K), is the absorption coefficient. Similarly, the portion of the energy traveling through the medium that is redirected or scattered is given as d] = - 0,1ds (3.2) SCG where or is the scattering coefficient for all directions. Since both absorption and scattering coefficients are linear, they may be combined and the resulting equation integrated once to yield the familiar Beer’s Law: 28 I(s)=I(0)e—‘6t’ (3.3) where ,6; = K; + 0'1 (3.4) is defrned as the extinction coefficient for the medium. For a particulate medium, this may be further expanded to fll = N (Csca + Cabs) (35) where N is the number of particles per unit volume; Cm and Cab, are the absorption and scattering cross sections which are a measure of the amount of extinction by a particle defined in terms of the ratio of the energy extinction rate to the incident irradiance. Beer’s Law gives an expression for the intensity of radiation at any points along the pathlength as a fraction of the intensity entering the medium at s = 0. The ratio I(s)/I(0) is the transmittance, and will be denoted here by the symbol :3. The subscript represents the case of transmittance for the particles-dispersed state. It is now left, before further development of a model to describe transmittance, to determine which mode of attenuation, if any, is dominant. In any medium, a beam of light is attenuated by both absorption and by scattering. Although absorption may be the dominant mode of attenuation, scattering is never entirely absent (Bohren and Huffman, 1983). Measurement of transmitted energy will therefore be influenced by a combination 29 of the effects of absorption and scattering. However if the efl‘ect of scattering is minimal, the medium can be modeled as one in which absorption is the sole cause of attenuation. For collections of particles, whether attenuation is dominated by absorption or not depends upon the size and optical properties of the particles. For the Zeolite particles, we first turn to a qualitative analysis. In the typical scattering analysis based on electromagnetic wave theory, particle surfaces are always assumed to be optically smooth. However, referring back to the electron-micrograph of Zeolites in Figures 2-5 will quickly show that such assumptions are strictly invalid since surface roughness levels, with voids on the order of microns, are large compared to the wavelengths of incident radiation studied here. Looking at a cross-section of the structure of a Zeolite particle, conceptually shown in Figure 3-1, one sees a network of grooves and cavities and is immediately reminded of surfaces that are designed to enhance absorption (Modest, 1993) by multiple reflections within the geometric boundaries of the surface. Zeolite Particle Figure 3-1 Cross Section of Zeolite Particle Showing Radiation Absorbing Cavities 30 Further observation of the Zeolite particles reveals yet another insight into their radiative properties. Since the surface characteristics indicate that absorption will be strong, we may be able to assume that the absorptive index k is reasonably large and is on the order of one. Considering the range of radiation wavelengths under study (500 nm S ,u S 800 nm), the particles being studied are of size (11 jam 5 2a 5 40 pm) that the particle size parameter used for scattering analysis x = 27mm (3.6) yields values that range from about 43 to 251. Since 10: >> 1, we may treat the particles as large opaque spheres and invoke the extinction paradox (van de Hulst, 1957), which states that a large particle will remove exactly twice the amount of light from an incident beam as it can intercept. Hence, the extinction efficiency, defined as Qext = Cext/ 7w) (3.7) where Cart = Cabs + Csca (38) will be always have a value of exactly two. Since particles of this size are described well by geometric optics, where the projected area of a particle for absorption and reflection is 71212 , half of this extinction efficiency is due to diffraction. Exactly how much of the other half is due to absorption, 31 and how much is due to reflection, will depend upon the reflectivity of the particle surface. This issue will be dealt with later. Working from the basic Beer’s Law and the definition of transmittance, equation (3.3) may be rewritten as rd =e'fl‘1‘ (39) where the pathlength variable s has been replaced with the total pathlength L of ER fluid, which is defined by the distance between the inner surfaces of the glass windows. Combining equations (3.5), (3.7), and (3.8) yields an expression for the extinction coemcient )6). = mzNQext (310) The number of particles per unit volume may be determined fiom the expression (3.11) 2 ll 5b where Vp is the volume of a single particle. If the particles can be modeled as spheres, this equation becomes N _ 3f, (3.12) 47ta3 32 Substituting into equations (3.9) and (3.10) yields a simple model for transmittance in terms of the ER fluid volume fraction, pathlength and particle size -3 L rd :8 ”/2. (3.13) The extinction coefficient may now be written as -211 (3.14) — 2a .51 Note here that the value of Qext = 2 has been substituted and that the denominator of the exponent is merely the particle diameter. 3.2 Experimental Results and Discussion Before analyzing transmittances for the purpose of making comparisons to the proposed model, we will return briefly to the issue of absorption as the dominant mode of attenuation. A clear indicator that absorption dominates the extinction process is a noticeable absence of the interference structure and the ripple structure in the extinction efliciency over a range of wavelengths (Bohren and Hufl'man, 1983). When this is the case, extinction and transmittance over that spectral range will be relatively constant and will not exhibit oscillations. This supports the most fundamental assumption for a perfectly 33 absorbing surface, that is, a black surface. In this idealized case, then there will be no wavelength dependence on transmittance. Figure 3-2 shows normal transmittance levels for ER fluids of different volume fractions in the particle-dispersed state, and clearly demonstrates the absence of large-scale oscillations as a function of wavelength (the visible variation in the plot is due to signal noise in the Spectrometer’s measurement). There is, however, a wavelength dependence for transmittance. As expected (Modest, 1993), we see that transmittance will increase with increasing wavelengths. 0.05 0-045 .- Dispersed particles - 0.04 .. zero field strength 0.035 c 0.03 0 0.025 ~~ 0.02 «L 0.015 -* °-°‘ ‘* fv = 0.0075 fv = 0.010 fv = 0.005 Transmittance §§§§§§§§§§§§§§§§ Radiation Wavelength Figure 3-2 Normal Transmittance Through ER Fluids of Different Volmne Fractions with Fully Dispersed Particles (L = 1.9mm, 2a = 11 um) Yet another indicator that scattering is less important can be made by the qualitative observation that scattering is a process that augments radiative intensity in directions other than that along the path of the incident beam. If scattering is to make a 34 significant contribution to the transfer process in the ER fluids studied, then transmittance should increase in directions away fiom that of the incident beam path. To test this, the light sensor was moved in small increments along a direction transverse to the incident beam and transmittance levels measured. The result is shown in the curve of Figure 3-3. After recalling that the diameter of the beam and sensor is 3 mm, and realizing that the total travel of the light sensor will be twice the diameter of the beam in order to travel the distance needed to fully cross the light beam, it is clearly seen that there is no transmittance in directions other than that of the beam. In this figure, the sensor position of zero represents perfect alignment with the source beam. CO CD 0 CID 00 W “Source-sensor 0'9 1 alignment ‘5 1 4- -> 0.8. Two Beam [j] . Diameters , - . 0-7 T .——Theoretical‘i 0'64; i :1 Measured I 0.5 ' _ 0.4 l 0.3 l Normallzed Transmittance “t-irr-i-i-H—t-Hrri-rfi- . '0. ‘9. “'2 “l ". “9 0’. ‘9. F. v.- 9 O x- N ('0 V’ II) Sensor Poeltlon (mm) Figure 3-3 Transnrittance as a Function of Sensor Position Transverse to the Axial Direction of the Incident Beam 35 Hence, from all of the evidence stated, our assertion is made that the dominant mode of attenuation in these ER fluids is absorption and that the model proposed (equation 3.13) will predict transmittance in the dispersed state. Prior to the presentation of transmittance data, uncertainty of measurements is addressed by considering equation (3.13) and rewriting transmittance in terms of transmitted energy through the ER fluid (0 divided by transmitted energy through an empty window (10). Measured energy transmitted through the ER fluid can be modeled as _3va (3.15) I = Ioe 2" For the PC 1000 spectrometer, intensity is measured as a voltage across an internal CCD camera in terms of analog-to-digital converter counts. Measurement error was then calculated by the formula: 2 2 2 + 0”] 5L 6L 0"! do 60 31 31, ,2 5f. + 51, 2 ”2 (3.16a) 5fv :l ERROR = [ Appendix C provides a detailed analysis of the sources of error. With reference to that analysis, it is seen that the error of equation (3.16a) is determined to be: ERROR = (o. 7 + 165.5 + 25.8 + 1231.2)”2 ~38 counts (3.16b) 36 We see here that the leading causes of error are the components due to measurement of volume fraction and particle size. The summation equation (3.16b) is divided by the initial, unattenuated incident radiation intensity of 1210 counts to obtain a 3% error. Again, this is detailed in Appendix C. Transmittance as a function of beam pathlength for ER fluids with volume fractions of 0.0025, 0.0050 and 0.0100 was first measured. Each of these fluids used particles that had an average diameter of 11 microns. From these measured transmittances for each volume fraction, equation (3.9) was used to determine corresponding extinction coefficients, and these coefficients were then averaged. Figure 3-4 shows the measured transmittance data points plotted on a natural logarithm scale with linear extinction lines drawn using the average extinction coefficients. 9 I—WL V”- -___ K a .33 l l l l l l l l .L. _ - _LAL. -1 a .._L..,L -.IL__ 8 _-~-——..--1<=1.30 Zerofield T f 7 «~ K323 r- strength ~—~- «e—T- — AA edeafleee—tfi ’5 a .L l "2.0025 . t. n3.-. Lajje A -_. AL ‘ 1v=.005 g 5.- 1. o1 .. +fma- v- - __ O =- i ‘ 1 . . 1 + a 4 .AV I : 7—5-19 1 ; .L a“ A$-A + A J_‘.- an .__ LIA £3 1 l l I eI : i . Fl L L. 1 t Ff -r-- ~~~~ ~—*---‘—+-- t—+ f—Tf 3' + ffiytr r. ”4.x, -%H T I 2 i ii. V 9' ' -‘t i l l i T . ‘7 T - T'FT ‘ T: g1: - T- T 1 _T _ I“; i i. 1A A .—_.—-_~_-:::f"f __,__,.. .L;L_.___. ._i.-:._,.e.-ea . TA; .1, -11, 1: 11 o N x. o. m. .- 01 v 0. «1 N 01 x. 0. «a n O O O O v- s- ‘- 1- N N N N Pathlengthtmm) Figure 3-4 Transnrittance in the Dispersed State for Varying Volume Fractions as a F rmction of Pathlength (2a = 11pm) 37 The experimental extinction coemcients of Figure 3-4 were determined by application of Beer’s Law (equation 3.3) to measured transmittances. Coefficients from multiple experiments were mathematically averaged to determine the reported coefficients. These were then compared to expected coefficients calculated directly from equation (3.14). The results are tabulated in Table 3-1. Volume fraction K [mm'l] [3 [mm'l] (experimental) (eq. 3.14) 0.0025 0.66 :l: 0.03 0.68 0.0050 1.39 :t 0.07 1.37 0.0100 2.80 :l: 0.38 2.73 Table 3-1 Experimental vs. Calculated Extinction Coefficients (Dispersed Particles, 2a = 11pm, L = 1.9 mm) Note that the largest error between experimental and calculated values is 2.6%. In further experimental work, transmittance through ER fluids made with particles of varying size was measured. As stated earlier, these particle diameters (2a) were 1 1, 25 and 40 microns respectively. The volume fraction for each of these fluids was held at a constant value of 0.005. Again using equation (3.9), average extinction coefficients are calculated based on these measurements. Figure 3-5 shows the measured transmittance data points with extinction lines drawn using the average extinction coefficients, much in the same way as Figure 3-4. 38 4511111141er “‘fll'l K-.37 l J i I I l “ r Zerofield 4 4e 4 l 3,5 4 ——————————— K859 ~ strength 44.-__4. .. .- -44 . 7.- ._ l A 4 K-1.42 I I A 4 4 3 3 4| . 4 . 4 4— -1.. _4_-_!4_._.l_.. _ 4 __ r: | ‘ 2831111!" I i i 1 E 4 g 2-5 *1 . 2a=25um g 4 tr f“: + {1+4 * rfl—H | 4 , ! 4 4 2 2 «4 . 2a=40um 4 ‘5 % vi -‘ l7 - ‘- E ‘T ‘T 4 [ T‘T"“ ' ' l 4 4 _4 t 1.5 ~ + 4 e+ ; , 4 4 +4 . _._ a. '_____,_-.4~, 4 c 4 4 i 4 I ' ..-~TI" I .4 4 . 4 4 4 4 44 4 4 I 1 ..__._,_+._ .1... ‘ +_ in ..,... __i_ _.,_ if A __ l fi+i¥ 71‘- J” --:;:._l _. s . 4 .. l 41* flwrwfi’” TT | A 4 ' 0.5 9* ’ :_ %i,;:_.;;f “W -: 1 0 r4 4 4 . 4 i , ; , o N. V. «a «a v- N V. 0. 0°. N N V. «9. «a n O O O O '- F '- P N N N N Pathlength (mm) Figure 3-5 Transmittance in the Dispersed State for Varying Particle Sizes as a Function of Pathlength (f; = 0. 005) Note here that the largest error between experimental and calculated values is 3.7%. Once again, these average extinction coefficients are compared to theoretical values calculated directly from equation (3.14), with these results tabulated in Table 3-2. Particle Diameter K [mm'l] [3 [mm'l] (experimental) (eq. 3.14) 11 um 1.42 :t 0.04 1.36 25 um 0.59 :l: 0.02 0.60 40 um 0.37 :t 0.05 0.38 Table 3-2 Experimental vs. Calculated Extinction Coefficients (Dispersed Particles, j; = 0. 005, L = 1.9 mm) All transmittance data is now collected and presented in Figure 3-6 for comparison to the fundamental Beer’s Law model of equation (3.9), confirming with 39 conclusive evidence that this basic model is highly capable of predicting radiative attenuation under the conditions of the investigation. 7T *4“ 4‘ 4‘ 4 4 4 4 T7 . . 4 4 . 6 ‘4‘ ‘T‘ Zero field “"1 4 fr 4 A ‘4 —~ -+:—,:4 I strength 4 4 I ' ~‘4 ’ 4 . 5 4‘ ——- —— 4 4 4 ._ - 4 .1-" .44 -In('l'ransmittance) Extlnctlon Coefflcle nt x Pathlength Figure 3-6 Comparison of Measured Transmittance to the Fundamental Beer’s Law (2a = 1 1m) Thus, we have established in this chapter that the dominant mode of radiant energy attenuation in an ER fluid made of Zeolite particles is absorption. A simple empirical model has been developed that describes attenuation of transmitted radiation energy while the particles are in a fully dispersed state. As can clearly be seen, the experimental data that was taken shows excellent agreement (errors less than 4%) between experimental levels of transmittance and those predicted fiom calculated extinction coefiicients using the absorption theory. This confirms that attenuation of radiant energy can be predicted by treating the dominant mode of extinction as absorption. 40 With this foundational work, it remains now to translate the same concept of attenuation by absorption to describe the effect of electric field-induced particle chaining, and its influence on the radiation transport properties. CHAPTER 4 - RADIATIVE TRANSFER THROUGH PARTICLES IN THE FIELD-INDUCED CHAINED STATE Since the Zeolite particles have been shown to be dominantly absorbing, it stands to reason that the cross-sectional area of the particle chains is the key to describing overall thermal energy attenuation for a beam of light of parallel incidence to the axis of the chains. Hence, the geometry of the particles at the interface between the glass electrode and the ER fluid (at the side of the window fi'om which radiation emanates) creates a series of “energy blockers”. It is left, therefore, to develop a simple model that quantifies the shape, area, and number of these blocking surfaces to provide for a way to model transmittance. 4.1 Additional Experimental Apparatus For this segment of the study it was desirable to observe chaining at a microscopic level in a direction transverse to the field lines. To accomplish this, an apparatus was devised in which two flat aluminum plates 0.25 millimeters thick were rigidly mounted with epoxy to a glass slide, as shown in Figure 4-1. These plates had progressive steps ground away from one edge, and were mounted so that a variety of distances (0.7 - 5.0 41 42 mm) separated them. Furthermore, provision was made on the aluminum plates for attaching wires from the high voltage amplifier, thus making them act as electrodes. ER fluid was placed in the gap separating the plates, and an electric field was applied to induce particle chaining. This apparatus was placed under an Olympus model BH-2 microscope with a Sony model XC-7ll CCD camera (pixel resolution 768 x 493) attached. Videotaped footage was obtained showing the transition of particles from the fully dispersed state to the field-induced chained state. The stepped-edge geometry of the aluminum plates allowed for simultaneous observation of the particle chains over multiple field strengths. ,—~— GLASS SLIP Gap Distances 0.7 - 5.0 mm v . ,_ ALUMINUM ELECTRODE Figure 4-1 Microscope Electrode Apparatus 4.2 Analysis of Particle Chain Geometry In the most ideal arrangement for the development of chains, all of the particles in the ER fluid will come together to form a series of uniformly spaced chains of single 43 particle width. This ideal line of single particles would stretch from one electrode surface to the other, as illustrated in Figure 4-2. Since the assumption is made that the spacing of these chains would be uniform, the total number of chains would then depend upon the total number of particles present and the number of particles needed to make one chain. That number is, of course, dependent upon the total pathlength or spacing between electrodes. In an idealized case, we also assume that all particles present in the ER fluid move to form chains. The total number of particles per unit volume (N) is given in terms of the volume fraction ([4,) and the volume of a single particle (VP) as (4.1) 2 ll 4‘3 Its \ \ ELECTRODES Figure 4-2 Uniformly Spaced Single-Particle Chains for the Ideal Case of Field-Induced Chaining 44 If the particles can be assumed to be modeled as spheres, then the total number of particles per unit volume is given by N 3 fv (4.2) 47:03 where a is the particle radius. The total number of particles (NT) in a given volume of ER fluid would then be found by simply multiplying the total number of particles per unit volume by the volume under consideration. To determine an expression for the volume of ER fluid, V, we consider a cylindrical beam of radiant energy that has unity cross-sectional area. As this beam passes through a slab of ER fluid, such as would be presented in the ER fluid filled window cavity, the total volume of ER fluid illuminated by this beam would be given in terms of the area of the cylinder face multiplied by the pathlength (L), or the distance separating the electrodes. The total number of particles present in this volume is now given by simple modification of equation (4.2): 3f.L (4.3) N, = 3 47m In a chain formed by single particles stretching over the entire pathlength, the total number of particles per chain is found by simply dividing the pathlength (L) by the 45 diameter of particles (20). Therefore, the total number of single particle chains possible in the cylindrical volume (C y) of irradiated ER fluid of unity cross-sectional area and pathlength L is determined by dividing the total number of particles in the volume (N T) by the number of particles per chain, which yields the expression: CV _ 6f‘, (4.4) 47m2 Now, in the ideal case, if each particle forming the single particle chain were perfectly aligned, the blocking area (A B) of each chain would simply be AB = m2 (4.5) With this, we have an expression for the total blocking area (A) of the ER fluid chains in the irradiated volume which is simply the blocking area of each chain (A B) multiplied by the total number of chains (Cy). Combining equations (4.4) and (4.5) yields the expression: A = 1.51:, (4.6) From this, it stands to reason that the energy transmitted through the volume of irradiated ER fluid (11.) will be attenuated by an amount that will be a function of this blocking area of the ER fluid chains. It was essentially this exact same concept that led to the basic model for transmittance for the dispersed-particle state (equation 3.13). The 46 model for chained-particle transmittance should therefore follow the same format, that is, an exponentially decaying function of blocking area and pathlength: —AL : 64.5va (4.7) We note at this point that particle size does not appear explicitly in this model. This is reconciled by realizing that volume fraction is a function of particle size. Also, the thesis behind development of this model was to compare the number of structures (particle chains) in a cross-sectional area to the amount of area void of structures. This, of course, is directly attributable to volume fraction. 4.3 Experimental Results and Discussion A careful study of the microscopic video footage taken of the chaining process revealed that the idealization of single particle chains, even in the earliest stages of chain formation, is far from reality. In fact, higher electrostatic forces at the electrode tend to pull particles from the chains to the interface, thus causing a “stacking” of particles into a conical geometry and widening the blocking area of each particle chain, as shown in Figure 4-3a. Figure 4-3b shows an actual microscopic video still showing this stacking effect. Visual analysis shows that this blocking area becomes as much as five particle diameters (10a) at the interface. This leads to a new expression for blocking area, similar to equation (4.5), where 47 AB = n(5a)2 = 2574442 (4-8) Multiplying this by the total number of chains per volume (equation 4.4) yields a new expression for total blocking area A = 37.5f. (4.9) and a new model for chained-particle transmittance = e—37.5f,L (4.10) \ STACKED PARTICLES ELECTRODE Figure 4-3a Stacking Effect of Particles at the Electrode Stacked Particles Electrode Surface Figure 4-3b Microscopic Video Still Showing Stacking Effect It has been shown in previous studies that transmittance through the field-induced particle chains of ER fluids increases as field strength increases (Zhang and Lloyd, 1992). Further analysis of video footage provides insight into the mechanism behind this phenomenon also. In this it was observed that particle chains that begin as typically uniformly spaced, single particle diameter structures begin to bend toward each other when field snengths are increased (conceptually shown in Figure 4-4a, video stills of actual phenomenon in Figure 4-4b). We note that as many as eight neighboring chains may be involved in this bending action around a single, central chain. This continues until the field is strong enough to make two or more of the particle chains connect, typically somewhere near their geometric middle at first (Figures 4-40 and 4-4d). But as time progresses, and/or field strengths increase, the connected region of the chains begins to increase (Figure 4-4c) until a new, single chain of multiple particle diameters is formed. 49 M \ ELECTRODES / Figure 4-4a Early Stages of Chain Bending Figure 4-4b Early Stages of Chain Bending 50 \ ELECTRODES / Figure 4-4c Initial Contact Between Chains Figure 4-4d Initial Contact Between Chains \ ELECTRODES Figure 4-4e Single Column of Multiple Chains Figure 4-4f Single Column of Multiple Chains 52 This suggests that a simple geometric model for transmittance as a function of field strength is appropriate, similar to that already developed in equation (4.10). Such a model would be constructed under the same assumptions as before, that is, transmittance would be determined by finding the total facial area blocked by particle structures. From that, it seems reasonable that by simply moving existing structures about within the same volume one would not expect very dramatic changes in transmittance, if any at all. The level of enhancement that does materialize may be a result of the tightening of particles held in the structure of the chain due to increased forces holding them together. All of this leads to the idea that the model already proposed (equation 4.10) may be sufficient to fully describe chained particle transmittance with the addition of an empirical modification, which is a function of field strength. Of course, the video footage can only provide evidence to support this contention. The validity of the argument will be determined from experimental data, and therefore determination of the empirical constant will be reserved until that time. Prior to presenting experimental results, uncertainty of measurements is addressed by considering equation (4.10) and rewriting transmittance in terms of transmitted energy through the ER fluid (1) divided by transmitted energy through a window filled only with silicone oil (1,). Measured energy can then be written as I ___ I e-37.5f,L (4.11) 0 with experimental uncertainty calculated by the formula: 53 2 + 2 + d 35L 0'! or 0761’ a: 0 2 )2 6f, ] (4.123) ERROR = [ Appendix C provides a detailed analysis of the sources of error. With reference to that analysis, it is seen that the error of equation (4. 12a) is determined to be: ERROR = (384.8 + 7150.4 + 495.2)”: z90 counts (4.12b) The magnitude of this level of error is in terms of the units of intensity as measured by the PC 1 000 spectrometer. We see here that the leading source of error is the component due to measurement of volume fraction. The summation equation (4. 12b) is divided by the initial, unattenuated incident radiation intensity of 1210 to obtain a percent error of 7.4%. In Figure 4-5 we see experimental transmittance data taken at Vf = 160 rms volts/mm for three difi‘erent sets of experiments with a variety of fluids of varying volume fraction. Superimposed against this is a line drawn to represent the model of equation (4.10). Table 4-1 summarizes comparisons of transmittance from experimental results to model prediction values. Note that the largest error between experimental and calculated values is 7.2%. 54 0'49 TT TT T T T‘ T 4»! _ LT; i 4 L; f //// 0.7 L ' LLL “ § 0 6 L .L L. L L L L // c o 4 r— 4 4 —— 4 [v 4// g 0 5 ' L—L »« - 4» L PJL—L . _4__a E I l f f—AL if ' LI L-——-—- Model b 03 L L . I” WM'T'I'L 9 #44 Ex erimental 5:- 02 L—L L L L44 L L l - p °‘ 4.2/”L *4 4 T ‘ 4 4 0 LLLLT444, ;; LLL O ‘- N ('3 If) h F ‘— N 888§8§8§§355 o o o o o c o o o o 0 Volume Fraction Figure 4-5 Transmittance in the Chained Particle State as a Function of Volume Fraction (Vf= I60 Vrms/mm, L = 1.9mm, 2a = 11 pm) Volume T (average 1, (eq. 4.10) fraction experimental) 0.0025 0.77 i 0.04 0.83 0.0050 0.69 :t 0.05 0.70 0.0075 0.59 :l: 0.07 0.59 0.0100 0.49 :t 0.08 0.53 0.0125 0.41 :t 0.03 0.43 Table 4-1 Experimental vs. Calculated Transmittance Levels (Vf= 160 Vrms/mm, L = 1.9 mm, 2a =11,um) Experimental evidence of the effect of field strength is shown in Figure 4-6. In this graphic, an ER fluid of volume fraction 0.0075 was subjected to increasing field strengths until arcing across the electrodes occurred. We observe that as field strengths 55 increase, the experimental variation tends to decrease. This supports the concept that greater order of the chain structure is achieved with greater field strengths. It also suggests that much more predictable transmittance levels will be obtainable with higher field strengths. Note that these voltage levels reported are RMS volts. It is also apparent that the relationship between field strength and transmittance is non-linear. However, the basic chained-particle model of equation (4.10) can easily be modified with a field strength based factor to accommodate changes in voltage levels and hence increase the robustness of the model. °'7"‘-" w ‘TT"4 I 4 l 4 l 0.5 .L . L . ,-L. t L. 4 . . L_T-;;:#IM91L. ‘ ' i "”7 4 0,5- 7. *’~- — , —— - i 4 0.4 .- __4 - 14-44;. 4—4 44—4 L44 4+ 4 -4 - . l 4 l L a 4 . I . 4 5 0.3 «-44-; 44; .41--. w L- .4444 “fi—e AL :7- _4.4 ‘ 4 _--—tlodol ‘ - ExportnontSott L 4‘f—‘T——fl ‘ Exporlmnt Set 2 4 w T 4'44 _ - -LJ- LU -. L 200 250 300 350 400 450 500 550 . 4 . Hold Strength (Vrm I I III III) Figure 4-6 Effect of Field Strength on Transmittance (I; = 0.0075, L = 1.9 mm, 20 =11 m) Figure 4-7 is provided to demonstrate the data correlation to a linearized model. Here, the natural logarithm of transmittance is plotted as a function of the inverse of the cube- root of field strength. -|n('|'ransml&ance) —t N -l 01 N at (a) .0 or [x F o‘ o‘ o' o‘ o‘ o‘ 6 Field Strength as a Function of tho Cube-root of Vrms/mm u—' s-' 0' Figure 4-7 Effect of Field Strength on Transmittance - Linearized Model ([1, = 0.0075, L = 1.9mm, 2a = 11 ,um) Based on the above data, an empirical model is proposed. First, two very broad assumptions: (1) At fields below 100 ms volts/mm the ability to drive particle chaining significantly drops off. This is evidenced in the significant drop in transmittance at that voltage level as seen in Figure 4-6, and (2) that no dramatic changes in the chaining process take place at fields higher than Vf = 500 rms volts/mm. The model proposed accounts for these assumptions very well. Also, when application of a higher field level was attempted, the insulating properties of the fluid broke down, causing arcing across the ER fluid, and the experiment was terminated. Regardless, based on the data taken, an enhancement to equation (4. 10) is proposed: ~37.5f,L46-253,——V 0 625] 2' = e / f+ ' (100 Vm/mmSVfSSOO Vrms/mm) (4.13) C 57 This model accounts for the effect of various field strengths on transmittance. Note that this model predicts that as field strengths approach zero the transmittance will decrease to that predicted by the dispersed-particle model. CHAPTER 5 - AN GULAR TRANSMITTAN CE ANALYSIS Until this point, all models have treated the problem of energy transmittance through an ER fluid composite window that is oriented so that its plane is held normal to the incident light beam (see Figure 5-1). As a result, in all cases of transmittance through particles held in the field-induced chained state, the particle chains have been aligned parallel to the energy path. Light Sensor Light Source Figure 5-1 ER Fluid Window with Normal Energy Incidence, Viewed fi'om the Top 58 59 In this chapter, we explore the case where the window is rotated about a vertical axis, as shown in Figure 5-2, thus presenting an oblique angle of incidence for the energy beam. The Zeolite particles are dominantly absorbing, which led us in Chapter four to the concept that the ends of the particle chains are the key to describing overall thermal energy attenuation for a beam of light of parallel incidence to the axis of the chains. In this case, however, the geometry of the entire particle chain, especially at the interface between both glass electrodes and the ER fluid will be responsible for creating the energy blocking objects within the medium. To quantify the level of transmittance will require only a modification to the existing model of Chapter four to provide for the angle dependent blocking area and pathlength. 5.1 Additional Experimental Apparatus Because of the angle sensitivity to transmittance measurements, it was clear that a very high degree of collimation of the energy source would be critical. As a result, the tungsten-halogen source used for the experiments in Chapters two and three was replaced with a helium-neon laser that not only provided collimated light, but was monochromatic at 632.8 nanometers. However, this created a problem in that the spectrometer became saturated with the higher intensity of light from the laser. To accommodate this, the spectrometer was replaced with an Oriel model 7070 photomultiplier tube (PM'I) and amplifier as the light sensor unit. 60 Additionally, the ER fluid window assembly and holder was mounted to a Zeiss Model 6334 rotary indexing table which allowed for precision control of rotation of the window about a vertical axis through the mass center of the ER fluid held in the orifice of the window. To insure precision and repeatability of mounting, locating pins were mounted on the Zeiss table and mating holes provided on the window mounting fixture. The result was provision for rotation of the window that would enable measurement of transmitted radiation as avfunction of azimuthal angle as is defined in Figure 5-2. Light Sensor \4 Light Source Figure 5-2 Rotation of ER Fluid Window About a Vertical Axis 61 A photograph of the experimental apparatus with the new light source / sensor equipment is given in Figure 5-3. Note that this is basically the same schematic of the apparatus shown in Figure 2-1. Prior to conducting experiments with the new light source / sensor equipment, calibration transmittances were measured through a sample of ER fluid and compared to levels measured with the spectrometer. Equivalent levels of measurement were achieved in all cases. Figure 5-3 Experimental Apparatus for Angular Transmittance Measurements 5.2 Analysis of Particle Chain Geometry at Oblique Angles In Chapter four, the incident light was held perpendicular to the plane of the ER fluid composite window. Because the electric field lines are also perpendicular to this plane, and the particles form chains that align with the field lines, the particle chains were 62 parallel to the incident energy’s path. This is shown schematically in Figure 5-4. As a result, based on the hypothesis of Chapter two in which the particles were found to be strong absorbers, and hence absorption was the dominant mode of energy attenuation, we needed only to consider the area of the ends of the particle chains to describe energy attenuation. 4 1 4 e {on ¢> e . 0 . ’fk.%’¢®96 @flflmefiw 5$$@03.% an 9.. Chain End Geometry Incident Energy Figure 5—4 Particle Chains Parallel to the Path of Incident Energy However, if the window is rotated about the vertical axis identified in Figure 5-2, two significant changes in the extinction problem occur. First, the profile of the blocking 63 area of the particle chain changes, as is shown in Figure 5-5. Secondly, the total pathlength of radiation travel through the ER fluid changes, which is also depicted in Figure 5-5. Both of these changes are of course directly attributable to the angle 0. Figure 5-5 Particle Chain Profile in a Rotated ER Fluid Window According to experimental evidence, the previously proposed models are accurate predictors of energy attenuation based on total blocking area due to particle geometries. Therefore, if modifications can be made that provide for the changes in pathlength and blocking area due to angle 6?, then the new models should provide satisfactory results for 64 transmittance through ER fluid windows held at oblique angles to the incident energy beam. With this, we begin development by first considering the change in pathlength. For an angle 0, there will be an obvious change in L associated with the cosine of 6. It can easily be seen that the efl‘ective pathlength of a particle chain presented to a beam of incident radiation at an oblique angle Bwill be given by the formula: Lefl=L/COSQ (5.1) We now consider the profile of the blocking area of the particle chain. As a chain is rotated some angle 0 about the vertical axis suggested in Figure 5-2, a significantly larger blocking area is presented. This area is composed of the particles forming the chain and the conical formation of particles on the window / particle interface opposite the direction of radiation incidence. This is illustrated conceptually in Figure 5-6, where the white particles represent those that are illuminated by, and hence absorbing, the incident radiation. 65 Figure 5-6 Blocking Area of Particles That are Exposed to Incident Energy It remains, therefore, to model the enhancement of blocking area presented by the additionally irradiated particles in order to model angle dependent transmittance. We begin to accomplish this by realizing that in three dimensions the rotated particle chain will present two distinct elliptical shapes that are connected by a line of particles. A conceptual representation of this is presented in Figure 5-7. Here, the primary ellipse represents the blocking area of the conical formation of particles closer to the source of incident radiation, and the secondary ellipse is associated with the particle formation farther away from the source. Between these two formations is a line of particles. However, for very small angles the line of particles will simply cast a shadow on the secondary ellipse, and will have no net effect on the overall blocking area. Another way 66 to think of this is to realize that whatever particles are in front of the elliptical blocking area will remove the same amount of radiation as the particles that their shadow covers on the conical formation. Secondary Ellipse '\ . . anaryEllrpse Incident Energy / / / Figure 5-7 Conceptualization of a Rotated Particle Chain in 3-d Consider now the aspect ratio between the effective diameter of the conical face (10 particle radii) and the pathlength. For the experiments reported here, this ratio is approximately 0.05. This leads to the conclusion that for angles greater than about three degrees the particle chains (between the conical particle formations) could add significant blocking surface to the secondary elliptical surface mentioned above. ' To further complicate the matter, we consider the formula for the number of chains per unit volume which was presented in Chapter four: 6f v (52) 67 For an ER fluid with a volume fraction of 0.0075, which was the average volume fraction used for this particular study’s experimental phase, this relationship provides for a maximum of about 118 single particle chains per square millimeter. It should be noted that a particle diameter of 11 microns is used here as well. Let’s assume then that there exists then about 118 columns per square millimeter. This means that there will be about 11 columns per linear millimeter, leading to the idea that the center-to-center distance between columns will be about 0.09 millimeters. Based on this, we consider the effect of turning the window and at what level we can expect columns to begin to “overlap” each other. Using a simple trigonometric analysis with a pathlength of 1.9 millimeters, it is found that overlap will start to occur at angles greater than about 2.75 degrees. This leads to the idea that since multiple overlaps will occur almost immediately upon turning the window we may model the blocking area presented by the rotated chains as a function of the effective diameters of the conical face. It has been previously shown that single particle chains are not necessarily what form upon introduction of an electric field. Indeed, with the fact that the particles form the conical geometries at the window interfaces and that particle columns may be composed of at least two single particle chains combined, a relationship between the true number of columns as a function of particle volume fraction must exist. However, prediction of this relationship analytically will be complicated by the fact that at lower volume fractions a larger number of “fewer particle chains ” may exist than at larger volume fractions. Similarly, for those larger volume fractions, where there are obviously 68 more particles available for development of multiple-chain columns, we may expect to find fewer total columns composed of very large numbers of particles. Therefore, treatment of this effect will be done empirically after experimental data has been collected. Based on these concepts, a model is proposed for blocking surface area. As noted above, the overlap of columns begins to occur at about the same angle that the chains between the conical formations begin to add blocking area to a single column. Therefore, we will combine these effects to model an angle dependent geometric blocking area as shown in Figure 5-8. A V Angle Sensitive Region Figure 5-8 Effective Blocking Area Thus, the particle columns now become modeled as cylindrical shapes as viewed from an oblique angle. After column overlap, these shapes will appear to extend across the entire area of irradiance on the window, and will be represented as a series of lines across the region (see Figure 5-9). 69 Figure 5-9 Overlapped Columns Extending Across the Entire Area of Irradiation The blocking area of this geometry is specified by the area presented by the ends of the conical faces at the glass interface, as has been previously developed, and an angular dependent component. The angular dependent portion will be composed of three terms. The first is a term describing the projected length of a single chain, which is simply the pathlength multiplied by the sine of the angle 0. The second term is the effective width of the chain, which is given by ten particle radii. Finally, the third term is the volmne fraction dependent total number of columns Cfl which will be determined empirically later. The product of these three terms provides the angle dependent component of blocking area, and for the proposed geometry the total blocking surface is given by A = 37541+ Cf" ~10a-1.9sin64 (5-3) 70 There remains still one more consideration before presenting a transmittance model. So far it has been assumed that the blocking area will be a function of the angle of incidence of the radiation beam with respect to the window. However, since the glass of the window and the air that the radiation travels through to reach it are different mediums with respect to their index of refraction, the angle of incidence and the angle of transmission are not the same. To determine this, we turn to Snell’s law which states sine, _1 (5.4) sinfl _ "2 where the subscripts represent the two different mediums and n is the index of refraction. For soda-lime glass and air, the indices of refraction are approximately 1.5 and 1, respectively. Ifwe treat 6?,- as the angle of transmission through glass, then it is related to the angle of incidence 0by: Q = Sin-1(sin&%2) (5.5) 4 f where, in this case n; is again the refractive index for air and n 2 is the refiactive index for ‘.‘-—- 4 the glass. This angle should be used for the blocking area model of equation (5.3), so that it becomes A = 375{1+ C, .10a -1.9sina} (5-6) 71 Similarly, equation (5.1) should be modified to utilize 6,- as well. With this we now provide a transmittance model that follows the same pattern as previously proposed models. In those models (see for example equation 4.7) transmittance was modeled as an exponentially decaying function of blocking area, volume fraction and pathlength. We may now combine equations (5.1) and (5.6) to propose a model for transmittance: 2' = e-4f.{37.5(1+C,v -10a-1.9sinq)}-%o s 62] (5.7) Here again, Cf, will be determined empirically after analysis of experimental data. 5.3 Experimental Results and Discussion Prior to the presentation of experimental results, uncertainty of measurements is addressed by considering equation (5.7) and rewriting transmittance in terms of transmitted energy through the ER fluid (0 divided by transmitted energy through an empty window (1,), as has been done in all previous analyses of measurement error. Measured energy transmitted through the ER fluid can now be modeled as I _ I e‘4fv {37.541+Cfv.10a-1.9sin6,)}.%080i] (5.8) 0 72 Measurement error was then calculated by the formula: 2 2 2 + 2 + 07 dado 0‘! 397““ a 07. d a a 670 +Z5L 0 + 2 R 4’. 4 (5.9a) ERROR = 4 Appendix C provides a detailed analysis of the sources of error. With reference to that analysis, it is seen that the error of equation (5.9a) is determined to be: ERROR = (858.6 + 508.4 + 79.2 + o + 18. 7)”? z38 counts (5.9b) The magnitude of this level of error is in terms of the units of intensity as measured by the Oriel PMT tube. We see here that the leading sources of error are the components due to measurement of incident energy intensity and volume fraction. The large variation in incident energy intensity is addressed later in this chapter. The summation of equation (5.9b) is divided by the initial, unattenuated incident radiation intensity which was 540 to obtain an uncertainty error of 7. 1%. Again, this is detailed in Appendix C. Transmittance as a function of window angle (off normal) for ER fluids with volume fractions of 0.0025, 0.0050, 0.0075, 0.0100 and 0.0125 was measured. Each of these fluids used particles that had an average diameter of 11 microns. From these measured transmittances for each volume fraction, an empirical relationship for ny was determined that provided for the best fit to data. Noting the similarity to the empirically determined field-strength term of equation (4.13), we have: 73 5 (5.10) Ur CI. = 3 >8 so that the transmittance equation becomes (5.11) -4}. {37541345745} .100.1_9sm014 4%05944 r=e Experimental data is shown graphically against model predictions in Figure 5-10. In this figure, note that the model lines represent different volume fractions, beginning with 0.0025 at the top and increasing in the order 0.0050, 0.0075, 0.0100 and 0.0125 as the lines progress downward. Furthermore, it should be noted that transmittance at 9 = 0 degrees has been adjusted to fit the normal transmittance model. This was done for two reasons. 4 4 Transmittance .-.-L-L- £,= 0.0100 Oflhcgk=piogshtfl°l~ oooooooooooooooooooooooo 4... TEL 4 4 Figure 5-10 Angular Transmittance: Experimental Data Versus Model Predictions (L = 1.9 mm, 2a =11,um, Vf=160 Vrms/mm) 74 First, during this phase of the investigation it was discovered that at ofi-normal angles (6 ¢ 0) there is an extremely strong dependence on angle for transmitted energy through an empty glass window (10). Consider for example Figure 5-11. Here we see an apparently periodic function for transmission of energy through glass only. Further, we see that the period seems to be on the order of about 0.3 degrees. Over the span of just a few one-hundredths of a degree, the level of transmission intensity can change as much as approximately 15 percent. Now this may initially suggest that intensity can be modeled as a sinusoidal firnction of angle and incident energy wavelength, since the apparent period is known fi'om data. But in another series of measurements, as shown in Figure 5-12, the apparent period has changed now to about 0.4 degrees, or about 33 percent from previous measurements given in Figure 5-11. Instead, this second set of data leads to another conclusion. 5‘ 5 E i s . 480 4’“ T ’ 4 T T'T4T 4% TT- -,L 4‘4 460«~— 4 4L _' L- 4 . L4- 8. 53. 8. 8. 9. 3. . E 8 to no to ID to ID ID ID ID Degree. from Normal Figure 5-11 Transmission Intensity of Energy Through Glass as a Function of Angle 75 Transmission Intensity Figure 5-12 Transmission Intensity of Energy Through Glass as a Function of Angie It is readily apparent that the data points collected may be samples of a much smaller wavelength signal. Indeed, using a simple math model that predicts the repeating of the incident energy signal’s intensity (wavelength = 632.8 run) every few thousandths of a degree, the data points collected and presented in Figure 5-12 are plotted against the incident energy waveform in Figure 5-13. From this, it is obvious that although the data points could be construed as representative of the incident energy waveform, they are more likely to be aliasing the true waveform. Now, to reverse this logic, we must consider what we have when we take data points of energy intensity without particles, which will be used in transmittance calculations. In essence, we have a level of energy that can change dramatically with the Pa... .;. .fl fun .In. 76 slightest change in angle (less than a few hundred-thousandths). It is not predictable, because we cannot be sure of the actual waveform it represents (equipment limitations prohibit further resolution of angle). Therefore, we must use a level of transmission that will be representative of all levels of possible intensities, that is, an average value. 380 360 340 320 300 280 260 240 Transmitted Intensity mvmmmnmvmmmom Qai'fiui‘fleiQm'YdQco'. co co co co co m e) N. ‘9 0’. to co co 3.75 3.85 3.95 Angle (degrees) Figure 5-13 Sampled Transmitted Energy Data Points Against a Model of the Waveform of Incident Energy Secondly, averaging the values of 1., provided for a more consistent form of transmittance in the data representation. Yet, it also provided a significant source of error in transmittance levels. However, the objective of this study was to identify and model the angular dependence on transmittance as previously defined. Therefore, to further negate the error from the uncertainty in 10 and shift the focus of the data more to the angular effects, the level of transmittance was adjusted to fit the model for normal transmittance by simply modifying 10 for each volume fraction. This modified 1., was then used to calculate transmittances for all data points at off-normal angles, providing the results shown previously in Figure 5-10. In essence, the data is forced to be correct to 77 the normal transmittance model so that the true profile of the angular effect can be considered. Because of the large amount of data, the results of the modified data values to the levels predicted by the model (equation 5.12) are presented in Appendix B. Corresponding measurement errors are presented there as well. Note that the average error is 3.44 percent, with the greatest error being 9.76 percent. CHAPTER 6 - OPTICAL SENSORS FOR STATE CLASSIFICATION The work of this investigation has demonstrated that ordered arrangements of the particulate in an ER fluid made of Zeolite particles causes significant and measureable changes in the level of radiation energy transmitted through a composite window. The levels of transmittance have been modeled for the cases of randomly dispersed particles (chapter 3), particles chained by application of an external electrical field (chapter 4), and for transmittance through chained particles at ofllnormal angles of incident radiation (chapter 5). During this study, it has been shown that transmittance is consistently predictable with less than ten percent error, and that these levels are functions of physical parameters of the ER fluid such as particle size, volume fraction, radiation pathlength and voltage field strength. A natural corollary to the results of this investigation is to consider the use of measured transmittance of radiant energy as a reliable means of qualifying the state of particle chaining. Clearly, a much broader ramification of this study lies in the idea that radiative transmittance may be used as a sensory quantity for use in control systems that can manage the level of particle chaining, and hence the unique properties of the ER fluid, for engineering applications. Consider the work of Radcliffe et al (1996) in which feedback control as an entirely new approach to ER fluid response management was investigated. The traditional 78 79 method of controlling ER fluids has been to simply apply an electric field and “see what you get” (that is, use open-loop control). If a precise relationship between a controlled system and a stimulus is known, this method of control can produce results that compare well to those desired. However, for an ER fluid, this relationship has yet to be fully established, especially in the case of repeated applications of an electric field. The use of feedback, or “closed-loop”, control overcomes the difficulties of open- loop control for systems such as ER fluids. Here, some form of measured system output is utilized by a controller to drive the system to a desired output. With this, the precise relationship between a controlled system and a stimulus is no longer needed, provided some form of accurate system state indication can be achieved. However, Andersland (1996) was quick to point out that “the missing element in previous work has been an accurate and precise sensor of ER fluid state”. Yet, in this foundational study, an optical sensor was used to indicate the state of the ER fluid and to provide feedback to a controller. The controller was intrinsically linked to the ER fluid stimulus, that is, the applied electric field. By defining a system error as the difference between an actual level of energy transmittance (the level of light illuminating the optical sensor) and a desired level of transmittance, the electric field was controlled to drive the ER fluid system toward some objective state. The results of this work were astounding. Experimental work showed that the use of the most simple proportional feedback control system yielded ER fluid response times (between a dispersed-particles to a chained-particles steady state) that were 35 times faster than those possible with open-loop control. Furthermore, the precision of actual 80 levels of transmittance as compared to desired levels was improved by a factor of 21 times. Hence, it was shown conclusively that simple feedback control systems could be employed to greatly enhance the management of an ER fluid system by overcoming slow, non-linear and non-repeatable ER fluid response. As an added benefit, this enhancement is achieved with an applied electric field that is more optimized, thus conserving system energy. This, of course, can become extremely important when considering remote applications. We now know that the objective state achieved by Andersland (1996) was an indicator of the particle chain geometry. Furthermore, as a result of the present work, mathematical models have been developed that link those geometries to levels of transmittance, as a function of the physical parameters of the ER fluid and the applied voltage field. Also, the models presented provide a way of predicting the level of transmittance enhancement expected with changes in the system. Specifically, knowing the efi‘ect of changes in field strength on transmittance, when giving consideration to the type of feedback control used by Andersland (1996), will provide for optimal control algorithms and hence optimal ER fluid systems. However, in the work of Andersland (1996) the optical sensor was not optimized. A silicon solar cell was used, along with a laser diode light source. This is not to suggest that this energy source / sensor pair was in any way inadequate, for it certainly achieved the results necessary to prove the concept. Yet, when considering ER fluid applications that are an integral part of some larger mechanical or thermal system, a significant level L. 81 of attention should be given to the sensor. The results of the present study provide some fundamental guidelines for this. In the present study energy transmittance models have been developed based on particle geometries, after showing that energy is removed by an irradiated particle by absorption. We have visually confirmed these geometries using a microscope, and have confirmed the validity of the models with experimental data. Integral to the development of these models has been the ability to quantify such parameters as the number of particle chains per unit area and the geometry of particle groupings at the electrode / particle chain interface. Linking these parameters to the physical parameters of the ER fluid particles (i.e. particle size, volume fraction, pathlength and angle of incidence) provides the foundation for the development of optical source / sensor pairs for such design constraints as size and functionality. 4: 4 41 ll CHAPTER 7 - CONCLUSIONS AND RECOMMENDATIONS FOR FURTHER STUDY This work has served to provide a comprehensive understanding of the physical problem of radiative heat transfer through an electrorheological fluid-based composite window. The underlying foundation of this work has been the assertion that for an ER fluid made from Zeolite particles, the dominant mode of energy attenuation is absorption. From this, analytical and empirical models have been derived that permit quantification of the levels of transmittance of incident energy through the ER fluids for a variety of particle states and physical parameters. These include analytical models for the case of transmittance through an ER fluid with its particles in a randomly dispersed state, and for transmittance through an ER fluid with its particles in an electric field-induced chained state. Also, an empirical enhancement to the chained-particle model was made to quantify the change in transmittance with changes in the level of applied electric field, and a semi-analytical / empirical model to determine the change in transmittance for the case of off-normal energy incidence was developed. Although the mode of heat transfer studied here was that of radiation, the physical understanding of the particle-particle interactions, for both a dispersed state and a field- induced chained state, will provide a foundation for studies that address conductive and convective heat transfer. In these future studies, the energy transport phenomena should 82 83 be readily quantifiable by considering once again the geometric forms of the particle structures within the ER fluid suspension. But a much broader ramification of this study, and hence the potential for a much greater contribution, lies in the area of utilizing radiative energy transmittance as a sensory quantity for determination of particle state. It is now well known that the particle chains are responsible for both changes in mechanical and thermal properties in ER fluids. Having the ability to simply and precisely sense the level of particle chaining with optical sensors will therefore have far reaching application in the control of both mechanical and thermal systems involving ER fluids. It takes little to imagine the possibilities of minimizing the size of optical source / sensor pairs. Sub-millimeter size light beams can encompass a number of particle chains to provide indication of the overall state of the ER fluid, and drive control systems that make applications viable and reliable. Of course, in the present study a “through-beam” source / sensor arrangement was utilized because of the availability of transparent electrodes. However, the use of reflective source / sensor pairs can open a wide variety of non-invasive sensing possibilities in systems where through-beam sensing is not possible. And then there is the potential for micron-sized source / sensors. As development of specific components progresses, such as micron-sized lasers, the feasibility exists to develop methods of classifying the activity of even smaller groups of particles yet, or perhaps even single particles. Future work may lead to understanding the dynamic response of particles as electric fields are being applied. That will allow prediction of upcoming particle geometries - before they even exist - and provide for the creation of 84 ultra-fast and ultra-precise control systems. We can hardly imagine what possibilities exist if we have the capability of literally tracking a single particle with a micron-sized source / sensor pair. Nor can we wait to find out. APPENDICES 84 ultra-fast and ultra-precise control systems. We can hardly imagine what possibilities exist if we have the capability of literally tracking a single particle with a micron-sized source / sensor pair. Nor can we wait to find out. a. a l I‘ i'ilil. 4|“. APPENDICES APPENDIX A AN OVERVIEW OF RECENT DEVELOPMENTS IN ER FLUID TECHNOLOGY The logical progression of ER fluids basic understanding, as outlined in Chapter 1, leads to new models of the behavior encountered. Many of these models came as a result of attempts to introduce ER fluids to commercial application. An example of this is the modeling of ER fluids when used in the oscillating squeeze-flow mode for a prototype automotive engine mount (Williams et al, 1993). Here, the authors report a solution for modeling the non-Newtonian behavior of ER fluids by introducing a “bi- viscous” characteristic. Another example of this was a study that considered distributed vibration control using ER fluids (Mahjoob et al, 1993). This work explored shortcomings in models for predicting mechanical properties of composite structural members filled with ER fluids. The authors present a method of extracting the “complex modules” of the ER fluids layer from modal parameters of the beam, and propose a modal modification scheme based on the discrete dynamic model of the composite beam. A study of ER non-Newtonian flow (Korobko and Mokeev, 1994) expanded existing models for mechanical behavior in low concentration suspensions to include ER fluids of high volume fiaction. The physical-mathematical model described here takes 85 if t .4 . 4 .13 86 into consideration the competition between the Maxwell tension in a disperse medium and the Maxwell pressure in a micro-nonuniform electric field between solid particles. Experimental data was presented that compares well to model predictions. An experimental investigation into the electrical modeling of ER fluids in the shear mode (I-Iosseini-Sianaki et al, 1994) studied the efl'ects of electrode separation and surface area, as well as fluid temperature, and the effect of these physical considerations on shear. This work also considered electrical characteristics of the ER fluids system, showing that capacitance does not vary significantly with field strength, shear rate or temperature, but does increase with increasing electrode surface area and reduces with increasing electrode separation. This study further showed that system resistance exhibits no reliable dependency on shear rate, but does indeed reduce with increasing field strength, temperature, and electrode surface area, while increasing with increased electrode separation. Historically there has always been an emphasis on potential applications for ER fluids that capitalize on the changes in mechanical properties associated with the ER effect. Recent work has included several studies that further the understanding of this mechanical behavior. For example, a recent theoretical and experimental study of ER fluids yield stress (Miyamoto and Ota, 1993) produced a calculation of static yield stress using a particle chain model. In this study, the Maxwell stress was calculated by direct solution of the Maxwell equation using a finite difi‘erential method. Creep behavior of ER fluids has also been the subject of a recent study (Otsubo and Edamura, 1994), which showed that creep recovery is purely plastic and the critical ‘1 t t .4 ‘1 l 87 stress corresponds to the static yield value. The authors of this work point out that although yield stress in steady shear can be predicted with the ideal chain model, instantaneous deformation without recovery below the yield stress cannot be predicted by this model. Further work has focused on creep and stress relaxation of ER fluids (Fisher, 1994). This investigation focused on characterizing the solid state properties of ER fluids in the pre-yield to post-yield stress region through dynamic mechanical analysis, stress relaxation and creep testing. Results showed that a range of viscoelastic properties exist, with dependence on temperature, applied stress mode, and the applied strain amplitude. Explanation of these results focused on the rearrangement of the particle structure induced by applied fields. As noted, with the evolution of the understanding of the fundamental mechanisms behind the ER fluid phenomenon, more and more effort is turned to the potential for commercial applications. A large portion of this work has been devoted to applications that seek to exploit the rheological properties of ER fluids for active vibration damping and control. A prime example of this was a study that dealt with vibration control of hollow cantilevered beams containing ER fluids (Choi et al, 1993). Here, a proof-of-concept experimental evaluation of the vibration properties of a hollow cantilevered beam filled with ER fluids was studied. The result was the development of an empirical model to predict field-dependent vibration characteristics. This was done by treating the beams as uniform viscoelastic materials and modeling as viscously-damped harmonic oscillators. In another vibration study on the effect of an ER fluids layer on a cantilevered beam 7:] ll 4 ll 88 (Haiquing et al, 1993), it was suggested that improved vibration control is gained by local application of ER fluids over that of a sandwich beam. In a similar study, the modeling and control of an ER fluids sandwich beam (Rahn and J oshi, 1994) was presented with advanced models for ER fluids structures based on viscoelasticity and sandwich beam theory. Simulations of a cantilevered ER fluids beam showed improved response stability and transient decay. A discussion of several applications for vibration control was provided by Morishita and Ura (1993). This work described four applications of active vibration control actuators and an adaptive neural-net control system suitable for controlling the actuators. Included were (1) an automotive shock absorber system, (2) a squeeze film damper bearing for rotational machines, (3) a dynamic damper for multi-degree-of- freedom structures and (4) a vibration isolator. Experimental support for vibrating beam control by a dynamic damper was provided. A further analysis of a short squeeze-film damper operating with ER fluids was provided by Jung and Choi (1995). From this work, a lubrication analysis of a short squeeze-film damper is carried out. The ER fluids were modeled as Bingham fluids for the development of the governing lubrication equation. With this, the authors reported a “significant increase in damping capability”. Another study involving the squeeze film damper for semi-active vibration control devices (Morishita et al, 1995), reported successful application of ER fluids to a squeeze film damper enabling a reduction in “whirling amplitude” in a flexible rotor. The model used in this study was based on the short bearing approximation of a journal bearing with 89 a Bingham plastic fluid being used. This work showed that the natural frequency of a flexible shaft was increased continuously as the applied electric field was strengthened, and that optimum damping for the flexible rotor for various rotating speeds could be achieved by controlling the field strength. In a similar study focusing on distributed damping of rotorcraft flexbeams using ER fluids (Kamath and Wereley, 1995), a model was developed to simulate the dynamic characteristics of ER fluids. The model uses two linear models for viscous and viscoelastic damping in combination with non-linear shape functions of strain rate. Model pararrreters are estimated for difi‘erent values of field strength using least squares. Damping and stiffness control in a mount structure using ER fluids (Kudallur et al, 1994) has also been studied. This work specifically addresses the design, construction and characterization of a mount structure for precision applications. Based on a one degree-of-fieedom system, the authors have shown that the ability to switch from an under-damped to an over-damped system at an appropriate switching time is better than critical damping in a situation where a mass is displaced fiom equilibrium and then released. The ER fluids in this work were modeled as a simple linear viscoelastic model (V oigt model with variable viscosity). Another damper study (Brennan et al, 1995) compared two types of ER fluids dampers. The findings of this work were that a damper using the shear mode of the fluid has a much greater range of damping than a damper operating in the valve mode. Again, the Bingham plastic model was used. The resulting non-linear damping model consisted of an off-field viscous damping component and an on-field Coulomb damping component. 90 In other works related to vibration damping and control, a proof-of-concept investigation on vibration control of a single-link flexible manipulator incorporating embedded ER fluids was carried out (Choi et al, 1995) This work placed an emphasis on simplified models and increased robustness in feedback control system. Also, an ER fluids damper for robots (Li et al, 1995) has been proposed and studied to control the elastic vibrations of robotics mechanisms. In this investigation, a device is described that works like a normal fluid damper when the ER fluids are not subjected to an electric field, yet provides great counter-torques when the field is applied. Several areas of applied research have been carried out that focus on the potential for automotive applications. As an example, a treatise on automotive engine mounts (Sproston et al, 1994) discusses a recently filed patent and highlights the potential advantages of using ER fluids in tension and compression rather than in shear. This followed a previous study on the potential for automotive engine mounts (Sproston et al, 1993), where a tenfold reduction in transmitted vibrations near resonance (4.25 Hz) was shown with fields of lOkV. Furthermore, an investigation into an anti-vibration mount utilizing ER fluids (Spurk and Muenzing, 1994), provided a theoretical description of the dynamic behavior of an ER fluids based engine mount under steady state harmonic excitation, with good experimental versus theoretical results shown. Another area of automotive application is that of enhanced suspension systems. An actively damped passenger car suspension system utilizing a low-voltage ER magnetic fluid has been described by Pinkos et a1 (1993). ER magnetic fluid technology is applied to a rotary shock absorber system with a control algorithm. Actual vehicle installation 91 with extensive testing was carried out, showing this to be a practical alternative to the currently used multi-damping shock absorbers. A similar study on the application of ER fluids to vibration control (Wei and Fu, 1994), discussed the potential for ER fluid-based shock absorbers. A new type of shock absorber. and control system was presented with test results. Finally, in an effort to expand ER fluid application beyond the confines of earth, a study on ER fluids for smart space materials was carried out (Lee et al, 1993). This work addresses the need for vibrational damping on large space systems characterized by lightly damped, large dimensional structures linked together at several joints, such as the multi-body truss structures used in space cranes. This work studied the use of ER fluid as a distributed actuator in flexible beam structures. In a similar work, a variable structure controller for a tentacle manipulator (Ivanescu and Stoian, 1995), suggested a new class of tentacle arms based on the use of flexible composite materials in conjunction with ER fluids. A model was proposed using Lagrange’s principle for infinite-dimensional systems, and was represented by a set of integral-difl‘erential equations. There has also been an abundance of non-vibrational control related applications studies performed in recent years. For example, a new type of linear piezoelectric stepper motor has been described (Dong et al, 1995) that combines the piezoelectric effect with the electrorheological effect. Here, the ER fluids act as a clamp while the piezoelectric actuator indexes the motor along a grid base. The authors cite advantages such as no vibration noise, zero wear, and high mechanical resolution. 92 ER fluids flow in journal bearings (Kollias and Dimarogonas, 1994) has also been studied to determine the effects of using ER fluids as lubricants in journal bearings. Here, the authors used the lubrication approximation of the equations of motion and conservation of mass. With this, a generalized non-linear Reynolds equation model based on the experimentally verified Bingham stress model was derived and solved by numerical techniques. Results showed an increased load capacity for journal bearings with increased electric field strength. A novel application of ER fluids for the dynamic performance enhancement of machine tool table systems has been described by Ishiyama et al, (1994). In this work, a method of reducing chatter vibration in rolling guideways during machining operations using an ER fluids-based variable damping mechanism was discussed. In a study of electrorheologically controlled landing gear, Lou et al (1993) discussed the application of ER fluids in aircraft landing gear. This work showed that energy absorption efficiency can reach almost 100% at various sink rates. And then a most interesting study on the development of ER fluids actuators for driving artificial muscles was performed by Tanaka and Tsuboe (1994). Here the authors described a method for obtaining an arbitral pressure and driving artificial muscle by controlling the electric field applied to an ER fluids. In an experimental apparatus described, it was confirmed that the artificial muscle can be driven smoothly for a variety of applied electric field types. APPENDIX B EXPERIMENTAL DATA FROM AN GULAR TRANSMITT AN CE MEASUREMENTS 4 Volume 4 4 Data . Model 4 4 Percent. 3 auction Angle , (9 (9 ’ Error. . 9 0.0025 -12 0.712 0.754 5.57 -10 0.733 0.767 4.43 -8 0.718 0.780 7.95 -6 0.762 0.794 4.03 .4 0.749 0.808 7.30 -2 0.764 0.822 7.06 2 0.767 0.822 6.69 4 0.788 0.808 2.48 6 0.788 0.794 0.76 8 0.764 0.780 2.05 10 0.733 0.767 4.43 12 0.728 0.754 3.45 0.0050 -12 0.597 0.595 0.34 -10 0.620 0.611 1.47 -8 0.631 0.627 0.63 -6 0.641 0.641 0.00 93 94 _ Volume ' : Data _ Model “:‘szmem. fiction ,. 441131.85 3 (9 * (9 - Err" 0.0050 .4 0.651 0.651 0.00 -2 0.654 0.681 3.97 2 0.690 0.681 1.32 4 0.677 0.662 2.27 6 0.672 0.645 4.19 8 0.612 0.627 2.39 10 0.605 0.611 0.98 12 0.613 0.595 3.03 0.0075 -12 0.489 0.474 3.17 -10 0.504 0.490 2.86 -8 0.518 0.507 2.17 -6 0.542 0.526 3.04 .4 0.569 0.545 4.40 -2 0.579 0.565 2.48 2 0.547 0.565 3.19 4 0.576 0.545 5.69 6 0.559 0.526 6.27 8 0.530 0.507 4.54 10 0.509 0.490 3.88 12 0.489 0.474 3.17 0.0100 -12 0.402 0.379 6.07 -10 0.418 0.395 5.82 -8 0.432 0.412 4.85 -6 0.429 0.430 0.23 .4 0.450 0.449 0.22 95 Percent Error 5.12 2.13 1.78 6.74 9.22 8.61 9.76 5.94 0.63 0.60 1.14 APPENDIX C EXPERINIENT AL UNCERTAINTY ANALYSIS Each segment of the investigation into radiative heat transfer through ER fluid windows has involved the development of models to predict the level of energy that is absorbed by the particulate medium in the ER fluid. The validity of these models was tested by gathering experimental data and making comparisons. In any experimental endeavor consideration must be given to measurement errors, those sources of variation that naturally skew the accuracy of experimental data. The effects of measurement error have been eluded to in each chapter, but are expanded upon and discussed in detail here. The methodology employed follows the guidelines set forth by Moflat (1988) for single- sample uncertainty analysis of experimental results. C.l Sources of Error There are many possible sources of error in an experimental system. For the apparatuses considered in this study, these included a series of measurement errors and bias errors. This includes inherent variations in the measurement devices, such as the spectrometer itself, which we had little control of. Along with this, we found that the 96 _ “1‘51 97 spectrometer itself, which we had little control of. Along with this, we found that the spectrometer had a problem in that it appeared to integrate a static signal over time. This caused the magnitude of the intensity of measurements to appear to drop, where in reality it was nearly a constant. Attempts to discover the source of this proved futile, and the manufacturer was unable to isolate this problem as well. As a result, this change in intensity proved to be one of the greatest sources of measurement error for the experiments. Other possible sources of error, that were beyond control, include fluctuations in the intensity of the light source resulting from changes in line voltage, and variations in the properties of the ITO coated glass used as electrodes. In the case of the glass, changes in the thickness of coating may cause a change in its reflectivity / transrnissivity characteristics. Changes in the thickness of the glass itself may also contribute to changes in these characteristics as well. Although no significant change in measured transmitted intensity was ever observed, this is nonetheless a possible source of error and should be noted. The most significant, and fortunately determinable, sources of error came from errors in making measurements of system parameters. These are summarized below: Variations in the measurement of ER fluid pathlength. As was noted in the chapter on experimental apparatus, the gasket material separating the glass electrodes was sulfur- cured styrene butadiene, which has rubber-like properties. As each window was placed in the window holding apparatus, the entire window assembly was sandwiched and held 98 together by tightening four finger screws. Changes in applied pressure against the gasket material created slight changes in its thickness, and thus the pathlength, and therefore are accounted for. Variations in the particle radius. The electron-photomicrographs of the Zeolite particles (reference Figures 2-5) clearly show that they are irregular in shape. Our approach to modeling these was to treat them as spheres of an average diameter. Determination of that average diameter came fiom measurement of the major and minor axes (perpendicular to each other) across a number of particles shown in each electron- photomicrograph. This was done by hand and compared to the scale printed on each picture by the electron microscope itself. Simple mathematical averages of these diameters (and corresponding radii) led to the determination of the average values used. Variations in particle volume fraction. Particle volume fraction was actually determined by creating ER fluids of known weight fiactions. Since the specific gravities of the Zeolite particles (1.10) and the silicone oil (1.11) are very nearly the same, the weight fi'action of particles to carrier fluid will effectively be the same as the volume fi'action, provided that no contaminating substances were present. Of these possible contaminants, only water was considered significant, especially since Zeolites are natural desiccants and adsorb water molecules immediately upon exposure to them. To eliminate adsorbed water molecules, the Zeolite particles were baked in an oven at 300°F for a period of time sufficient to “boil off” any attached water molecules. To create the desired weight fraction, exact amounts of silicone oil and Zeolite particles (measured with a Mettler electronic scale) were mixed immediately upon removal of the baked Zeolite particles 99 from the oven. However, it must be remembered that the Zeolites again began adsorbing water at the instant of exposure. In this case, the transfer from oven to silicone oil meant exposure to air. Although the ambient humidity was low, some water did indeed reattach in this brief period of time, causing the weight of the Zeolite to increase, and hence imparting error on the volume fi'action measurement. Variations in field strength. Field strength was determined by setting the output of a function generator to a voltage level between 0 and 2000 millivolts, and then inputting that signal to a high voltage amplifier that had a fixed gain of 1000. The amplitude of the function generator’s output was determined by visually measuring the signal on an oscilloscope, which imparted a certain level of error. Variations in window angle. As was discussed thoroughly in chapter 5, a slight change in window angle caused considerably different levels of transmitted intensity to be measured. A micrometer scale on the Zeiss rotary table was used to provide positioning with accuracy to the nearest 1/60th of a degree. However, as was demonstrated, much smaller variations in angle resulted in difl‘erent transmitted energy measurements. Therefore, a significant source of error was presented. Furthermore, because the window holding apparatus was designed so that the window could be removed for experiment preparation, a certain level of positioning variation existed, especially as bearing materials began to wear over time. lOO C.2 Error in the Dispersed Particles Model Recall from chapter 3 the dispersed particles model: _344 (3.15) I=Ioe 2" Using the chain rule of differentiation leads to 1 2 + 07 E6“ 07 E51. 2 0'! +—v+ it? 076% 0 d ERROR = I: ,.8 ] (3.16a) where the 5 parameter represents the variation observed for each specific measurement, and: J _3ivL (C.l) _ = e a 5?; 3f,L _ C2 1: I e— 2a ( 3L) ( ) 2. ° 2a 3va . 51:] e_ 20 (_3fVJ (C3) OI ° 2a 101 o7 -ifv—L (- 3 f, L) (CA) 2a2 For the series of experiments carried out to measure transmitted intensity in an ER fluid with its particles in a randomly dispersed state, the following parameter values and variations may be considered to be typical: Parameter Value Variation Initial Intensity (10) 1210 40 Volume Fraction (fv) 0.0075 0.001 Pathlength (L) 1.9 (mm) 0.1 Particle Radius (a) 5.5 x 10'3 (mm) 2 x 10'3 Using these, and applying to equations (C.l) - (C4) and (3.16a) yields an expected error level in measured intensity of 37.7 analog-to-digital converter counts. In terms of transmittance, after dividing by I, = 1210, this leads to an expected error level of i 3%. C3 Error in the Chained Particle Model Recall from chapter 4 the chained particles model: I = I e-mm (4.11) 0 Again using the chain rule of differentiation leads to a statement of the expected error, which is given as the change in measured intensity I, or 102 2 + 2 + 07 Z90 a a 21—51, 56L T (4.12a) The 6 parameter, as before, represents the variation observed for each specific ERROR = I: measurement, and: 1: €37.5va (C.5) 07.. (C.6) g = [Gem's/"L (_ 37.51.) (0.7) 37 5 _fl _ — . f, L _ Ice (— 37.5 f,) For the series of experiments carried out to measure transmitted intensity in an ER fluid with its particles in a field-induced chained state, the following parameter values and variations may be considered to be typical: Parameter Value Variation Initial Intensity (10) 1210 40 Volume Fraction (fv) 0.0100 0.002 Pathlength (L) 1.9 (mm) 0.1 103 Using these, and applying to equations (C.5) - (C7) and (4.12a) yields an expected error level in measured intensity of 89.6 counts. In terms of transmittance, afler dividing by I, = 1210, this translates to an expected error level of i 7.4%. C.4 Error in the Chained Particle Model With Field Strength as a Parameter The chained particle model of chapter 4 was modified to include an empirical term to account for the change in transmittance as a function of field strength. A model was proposed with field strength as a parameter: (4.13) -37.5 1,497 ] s/V, +0.625 r=e C which is rewritten here in terms of measured intensity: I = I e-375 1.49% V, +0625] (Q8) 0 From this, an equation for error can be written: 2 + 2 + 2 l W! 1 41 a a d 07, Ea. .1}. <7. 5"; 2 a ] (C.9) ERROR = [ 104 where, as before, the 6' parameter represents the variation observed for each specific measurement, and: a -37.5f,L[6-7W+0‘625] (C. 10) at" , 925,, (C11) 51 =Ie 3 5fL[ /l/+0.625] '[-375L[ 6.25 ]] 9‘. ° {/2 + 0.625 07 "37'sf'L[6'% V, +0.62 J l: [ 625 II (C' 12) —=I,e - —375f, — 01 3M + 0.625 (0.13) ’37" 'L 625w +0625 1:10,, ’1/J7 1.137% 625 2 5V! 3 (fo + 0.625) V25 For the series of experiments carried out to measure the effect of field strength on transmitted intensity in an ER fluid with its particles in a field-induced chained state, the following parameter values and variations may be considered to be typical: Parameter Value Variation Initial Intensity ([0) 1210 40 Volume Fraction (fv) 0.0100 0.002 Pathlength (L) 1.9 (mm) 0.1 Field Strength (V f) 200 (V rmts/mm) 20 105 Using these, and applying to equations (C.10) - (C.13) and (Q8) yields an expected error level in measured intensity of 89.6 counts. In terms of transmittance, after dividing by I, = 1210, this translates to an expected error level of i 7.4%. C.5 Error in the Angular Transmittance Model The angular transmittance model of chapter 5 was developed to quantify transmittance of chained particles when the ER fluid window is rotated some angle off normal about a vertical axis through the ER fluid mass center. A model was proposed that included an empirical term to provide for a volume fraction dependent total number of particle columns: (5.11) —[f, {37.5[1+5%/—£-106-1.9sina) }%osa] r=e It should also be noted again that 6) is the transmittance angle of light through the soda-lime glass. It is determined fiom Snell’s Law (equation 5.4), and is a fimction of the measured incidence angle 6?. From this, an equation for error can be written: 2 + 2 2 d 506a or 9‘. o”! 0161. + 2 a +550, + a O a; O a 2 % (5.9a) Em=[ a ] 106 where once again the 6 parameter represents the variation observed for each specific measurement. To perform the necessary partial differentiations, the exponential term is written out in a full term-by-term expansion (note that the transmittance equation is also written now in terms of intensities): 2 {37.5 f,L/coso,. +3918.75 Manna/66319] (C-14) I=Ie 0 We may also write equation (C. 14) as: I = I e-fl (C.15) where ,B = 375va/cosfl, + 3918.75 flame, (016) From this we can now write the partial derivatives: 0‘1 _fi (019) 2,70- = e (021) g = Ioe’” -[375L /cos€,. +26125f;%aLtan9,] V 107 c.22 % = Ice—p '[375f, /cos6l, +3918.75f,%atan6,] ( ) a ‘ (C23) —_ 5 2/3 a: —1,e -[3918.75f, LtanQ] fl _ -3 375/,Lsino/ % , ] ((3.24) 50f!” .[ icoszgi+391875fv aLsec o, For the series of experiments carried out to measure the effect of off-normal incidence on transmitted intensity in an ER fluid with its particles in a field-induced chained state, the following parameter values and variations may be considered to be typical: Parameter Value Variation Initial Intensity (lo) 540 50 Volume Fraction (fv) 0.0075 0.001 Pathlength (L) 1.9 (mm) 0.1 Particle Radius (a) 5.5 x 10'3 (mm) 2 x 10'3 Off-normal Angle (9,) 0 (degrees) 0.0167 Using these, and applying to equations (C.20) - (C.24) and (C.14) yields an expected error level in measured intensity of 38.3 counts. In terms of transmittance, after dividing by I, = 540, this translates to an expected error level of :l: 7.1%. In addition to the estimation of uncertainty in the overall result, confidence intervals around the mean extinction coeficients of chapter three are considered as a final indicator of the integrity of these parameters. For example, Table 3-1 reports an «I— 108 extinction coefficient calculated from equation (3. 14) of ,B = 0.68 for a volume fraction of 0.0025. Data points collected at this volume fraction yielded experimental extinction coeflicients of 0.64, 0.64, 0.65 and 0.69. Based on these, a confidence interval is constructed using the t-distribution with three degrees of freedom and an estimate of the standard deviation based on the equation: 2 n§Xi2_(;XJZ (C22) S = n(n-1) In this case, a 95% confidence interval for the true value of the coefficient is found to be: 06zsps070 Computations for the remaining extinction coemcients produce similar results, which leads to the conclusion that the reported values are indeed the best estimate for the result, and, with 95% confidence, the true value is believed to be within the limitations set forth by the percent error for each experiment. BIBLIOGRAPHY BIBLIOGRAPHY Adolf, D., and Garino, T., 1995, “Time Dependent Dielectric Response of Quiescent Electrorheological Fluids,” Langmuir, Vol. 11, No. 1, pp. 307-312. Adolf, D., Garino, T., and Hance, B., 1995, “Permittivity of Electrorheological Fluids Under Steady and Oscillatory Shear,” Langmuir, Vol. 11, No. 1, pp. 313-317. Andersland, R. M., 1996, “Feedback Control of Electrorheological Fluids,” Master ’s Thesis, Michigan State University, East Lansing, Michigan. Bohren, C. F., and Huffman, D. R., 1983, “A bsorption and Scattering of Light by Small Particles. ” Brennan, M. J., Day, M. J., and Randall, R. J., 1995, “Electrorheological Fluid Vibration Damper,” Smart Materials and Structures, Vol. 4, No. 2, pp. 83-92. Brewster, M. Q., and Tien, C. L., 1982, “Radiative Transfer in Packed Fluidized Beds: Dependent Versus Independent Scattering,” J. of Heat Transfer 104, 573. Buckius, R. 0., 1986, “Radiative Heat Transfer in Scattering Media: Real Property Considerations,” Proc. of Eighth International Heat Transfer Conf., Vol. 1, 141. Ceccio, S. L., and Wineman, A. S., 1994, “Influence of Orientation of Electric Field on Shear Flow of Electrorheological Fluids”, Journal of Rheology, Vol. 38, No. 3, pp. 453- 463.. Choi, S., Park, Y., and J ., 1993, “Vibration Characteristics of Hollow Cantilevered Beams Containing an Electrorheological Fluid,” International Journal of Mechanical Sciences, Vol. 35, No. 9, pp. 757-768. Choi, S., Thompson, B. S., and Gandhi, M. V., 1995, “Experimental Control of a Single- Link Flexible Arm Incorporating Electrorheological Fluids,” Journal of Guidance, Control and Dynamics, Vol. 18, No. 4, pp. 916-919. Conrad, H., Shih, Y., Chen, Y., 1994, “Effect of Prior Heating Zeolite / Silicone Oil Suspensions on Their Subsequent Electrorheology at 25 Degrees C,” Developments in 109 110 Electrorheological F lows and Measurement Uncertainty - 1994 American Society of Mechanical Engineers Fluids Engineering Division, Vol. 205, pp. 83-87. Dong, S., Li, L., Gui, Z., Zhou, T., and Zhang, X., 1995, “New Type of Linear Piezoelectric Stepper Motor,” IEEE Transactions on Components, Packaging and Manufacturing Technology - Part A, Vol. 18, No. 2, pp. 257-260. Fisher, M. D., Sprecher, A. F ., and Conrad, H., 1993, “Metallography Applied to the Structure of Electrorheological (ER) Fluids,” Metallography: Past, Present and Future ASTM Special Technical Publication, No. 1165, pp. 372-3 89. Fisher, M. D., 1994, “Creep and Stress Relaxation of ER Fluids,” Developments in Electrorheological F lows and Measurement Uncertainty - 1994 American Society of Mechanical Engineers Fluids Engineering Division, Vol. 205, pp. 89-98. Ginder, J. M., 1993, “Difliuse Optical Probes of Particle Motion and Structure Formation in an Electrorheological Fluid,” Physical Review, Vol. 47, No. 5, pp. 3418-3429. Ginder, J. M., and Ceccio, S. L., 1995, “Effect of Electrical Transients on the Shear Stresses in Electrorheological Fluids,” Journal of Rheology, Vol. 39, No. 1, pp. 211-234. Haiquing, G., King, L. M., and Cher, T. B.,1993, “Influence of Locally Applied Electrorheological Fluid Layer on Vibration of a Simple Cantilever Beam,” Journal of Intelligent Material Systems and Structures, Vol. 4, No. 3, pp. 379-3 84. Halsey, T. C., 1993, “Coarsening of a Quiescent Electrorheological Fluid 11: Theory,” Materials Research Society Symposium Proceedings, Vol. 289, pp. 57-61. Hosseini-Sianaki, A., Bullough, W. A., Tozer, R. C., Whittle, M., and Makin, J., 1994, “Experimental Investigation into the electrical modeling of Electrorheological Fluids in the Shear Mode,” IEEE Proceedings: Science, Measurement and Technology, Vol. 141, pp. 531-537. Hunter, L. W., Mark, F. F., Kitchin, D. A., Feinstein, M. R., Blum, N. A., Platte, B. R., Arcella, F. G., Kuespert, D. R., and Donohue, M. D., 1993, “Optical Effects of Electrorheological Fluids,” Journal of Intelligent Material Systems and Structures, Vol. 4, No. 3, pp. 415-418. Ishiyama, M., Aoyama, T., and Edamura, K., 1994, “Application of Electrorheological Fluid for the Dynamic Performance Enhancement of Machine Tool Table Systems (Analysis of the Dynamic Characteristicsof a Simple Table System),” Transactions of the Japan Society of Mechanical Engineers - Part C, Vol. 60, No. 577, pp. 2964-2970. 111 Ivanescu, M., and Stoian, V., 1995, “Variable Structure Controller for a Tentacle Manipulator,” Proceedings - IEEE International Conference on Robotics and Automation, Vol. 3, pp. 3155-3160. Jung, S. Y., and Choi, S., 1995, “Analysis of a Short Squeeze-film Damper Operating with Electrorheological Fluids,” T ribology Transactions, Vol. 38, No. 4, pp. 857-862. Kamath, G. M., and Wereley, N. M., 1995, “Distributed Damping of Rotorcrafi F lexbeams Using Electrorheological Fluids,” AIAA / ASME / ASCE / AHS Structures - Structural Dynamics & Materials Conference - Collection of Technical Papers, Vol. 4, pp. 2297-2307. Kanu, R. C., and Shaw, M. T., 1994, “Eflects of AC and DC Electric Fields on the Dynamic Shear Response of Electrorheological Fluids Consisting of Rod-like Particles,” Developments in Electrorheological F lows and Measurement Uncertainty - I 994 American Society of Mechanical Engineers Fluids Engineering Division, Vol. 205, pp. 1 17-124. Kerker, M., 1969, “The Scattering of Light and Other Electromagnetic Radiation, ” Academic Press, New York. Kimura, H., Minagawa, K., and Koyama, K., 1994, “Induced Network Structure in Liquid Crystalline Polymer Evidenced from Electrorheological Normal Stress Measurements,” Polymer Journal, Vol.26, No. 12, pp. 1402-1404. Kita, Y., Ohshima, T., Takase, H., Kondoo, I., and Fujisawa, S., 1993, “Fundamental Study of Electrorheological Fluids,” Transactions of the Japan Society of Mechanical Engineers, Part B, Vol. 59, No. 562, pp. 1816-1821. Klingenberg, D. J., van Swol, F. and Zukoski, C. F., 1989, “Dynamic Simulation of Electrorheological Suspensions,” Proceedings of the ACS Division of Polymeric Materials Science and Engineering, Vol. 61, pp. 154-155. Kollias, A., and Dimarogonas, A. D., 1994, “Electrorheological Fluid Flow in Partial Journal Bearings,” Developments in Electrorheological Flows and Measurement Uncertainty - I 994 American Society of Mechanical Engineers Fluids Engineering Division, Vol. 205, pp. 69-81. Korobko, E. V., and Mokeev, A. A., 1994, “Study of Electrorheological Non-Newtonian Flow,” Developments in Electrorheological F lows and Measurement Uncertainty - 1994 American Society of Mechanical Engineers Fluids Engineering Division, Vol. 205, pp. 147-154. Kudallur, S. B., Conners, G. H., and Bufl'mgton, K. W., 1994, “Damping and Stifiiress Control in a Mount Using Electrorheological (ER) Fluids,” Developments in 112 Electrorheological F lows and Measurement Uncertainty - 1994 American Society of Mechanical Engineers Fluids Engineering Division, Vol. 205, pp. 57-61. Lee, G. K., Yuan, F., and Conrad, H., 1993, “On the Use of Electrorheological Fluids for Smart Space Materials,” Space Exploration Science and Technologies Research - American Society of Mechanical Engineers Aeorspace Division, Vol. 31, pp. 107-115. Li, J ., Jin, D., Zhang, X., Zhang, J ., and Gruver, W. A., 1995, “Electrorheological Fluid Damper for Robots,” Proceedings - IEEE International Conference on Robotics and Automation, Vol. 3, pp. 2631-2636. Lou, Z., Ervin, R. D., Filisko, F. B., and Winkler, C. B., 1993, “Electrorheologically Controlled Landing Gear,” Aerospace Engineering, Vol. 13, No. 6, pp. 17-22. Mahjoob, M. J ., Martin, H. R., and Ismail, F., 1993, “Distributed Vibration Control Using Electrorheological Fluids,” Proceedings of SPIE - The International Society for Optical Engineering, Vol. 1923, part 1, pp. 768-774.. Martin, J. E., Odinek, J., and Halsey, T. C., 1993, “Coarsening of a Quiescent Electrorheological Fluid 1: Experiment,” Materials Research Society Symposium Proceedings, Vol. 289, pp. 49-56. Martin, J. B., and Odinek, J ., 1995, “Light Scattering Studies of an Electrorheological Fluid in Oscillatory Shear,” Dynamics in Small Confining Systems 11 - Materials Research Society Symposium Proceedings, Vol. 366, pp. 155-162. Miyamoto, T., and Ota, M., 1993, “Theoretical and Experimental Study on ERF ’s Yield Stress”, Transactions of the Japan Society of Mechanical Engineers, Part B, Vol. 59, No. 562, pp. 1822-1828. Modest, M. F ., 1993, “Radiative Heat Transfer, ” McGraw-Hill, New York. Mofi‘at, R. J., 1988, “Describing the Uncertainties in Experimental Results,” Experimental Thermal and Fluid Science, Vol. 1, pp. 3-17. Morishita, S., and Ura, T., 1993, “BR Fluid Applications to Vibration Control Devices and an Adaptive Neural-Net Controller,” Journal of Intelligent Material Systems and Structures, Vol. 4, No. 3, pp. 366-372. Mosishita, S., An, Y. K., and Mitsui, J., 1995, “ ER Fluid Squeeze Damper for Semi- active Vibration Control Devices: A Parametric Study,” ASME Paper 95-GT-251. Otsubo, Y., and Edamura, K., 1994, “ Creep Behavior of Electrorheological Fluids,” Journal of Rheology, Vol. 38, No. 6, pp. 1721-1733. 113 Pinkos, A., Shtarkman, E., and Fitzgerald, T., 1993, “Actively Damped Passenger Car Suspension System with Low Voltage Electrorheological Magnetic Fluid,” Vehicle Suspension and Steering Systems - SAE Special Publications, No. 952, pp. 87-93. Radcliffe, C. J., Lloyd, J. R., Andersland, R. M., and Hargrove, J. B., 1996, “State Feedback Control of Electrorheological Fluids”, ASME International Mechanical Engineering Congress and Exhibition and ASME J. Dynamic Systems - Measurement and Control, Atlanta, GA, November 1996. Rahn, C. D., and Joshi, S., 1994, “Modeling and Control of an Electrorheological Sandwich Beam,” Active Control of Vibration and Noise - American Society of Mechanical Engineers Design Engineering Division, Vol. 75, pp. 159-167. See, H., Tamura, H., Doi, M., 1993, “Role of Capillary Forces in Electrorheological Fluids,” Journal of Physics D: Applied Physics, Vol. 26, No. 5, pp. 746-752. Shu’lman, Z. P., 1982, “Utilization of Electric and Magnetic Fields for Control of Heat and Mass Transfer in Dispersed Systems (Suspensions),” Heat Transfer - Soviet Research 14(5), 1, September - October. Sproston, J. L., Stanway, R., Prendergast, M. J., Case, J. R., and Wilne, C. B., 1993, “Prototype Automotive Engine Mount Using Electrorheological Fluids,” Journal of Intelligent Material Systems and Structures, Vol. 4, No. 3, pp. 418-419. Sproston, J. L., Stanway, R., Williams, E. W., and Rigby, S., 1994, “Electrorheological Automotive Engine Mount,” Journal of Electrostatics, Vol. 32, No. 3, pp. 253-259. Spurk, J. H., and Muenzing, R., 1994, “Anti-vibration Mount with ER-fluid,” Developments in Electrorheological F lows and Measurement Uncertainty - I 994 American Society of Mechanical Engineers Fluids Engineering Division, Vol. 205, pp. 51-56. Tabatabai, S., 1993, “Aspects of Radiation Heat Transfer in ER Fluid Based Composite Windows,” Studienarheit, Rheinisch-Westfalrsche Technische Hochschule Aachen / Michigan State University, 84 pages. Tanaka, Y., and Tsuboe, H., 1994, “Development of ERF-actuator for Driving Artificial Muscle,” Transactions of the Japan Society of Mechanical Engineers - Part C, Vol. 60, No. 579, pp. 3882-3887. Tao, R., and Sun, J. M., 1991, “Three-Dimensional Structure of Induced Electrorheological Solid,” Physical Review Letters, July 15, Vol. 67, No. 3, pp. 398-401. Tsukiji, T., and Furuse, N., 1994, “Flow Characteristics of ER Fluids Between Concentric Cylinders,” Developments in Electrorheological F lows and Measurement 114 Uncertainty - 1994 American Society of Mechanical Engineers Fluids Engineering Division, Vol. 205, pp. 7-14. van de Hulst, H. C., 1957, “Light Scattering by Small Particles, ” Wiley. Wei, C., and Fu, Z.., 1994, “Electrorheological Fluid and Its Application in Vibration Control,” Journal of Beijing Institute of Technology, Vol. 3, No. 1, pp. 91-98. Williams, E. W., Rigby, S. G., Sproston, J. L., and Stanway, R., 1993, “Electrorheological Fluids Applied to an Automotive Engine Mount,” Journal of Non- Newtonian Fluid Mechanics, Vol. 47, pp. 221-238. Winslow, W. M., 1947, U. S. Patent 2,417,850. Zhang, C., and Lloyd, J. R., 1992, “Measurements of Radiation Heat Transfer in Electrorheological Fluid Based Composite Materials,” Developments in Radiative Heat Transfer, HT D, Vol. 20, 1992 National Heat Transfer Conference, San Diego, CA, Aug. 8-11, pp. 55-62. Zhang, C., and Lloyd, J. R., 1993, “Enhancement of Conductive Heat Transfer Through an Electrorheological Fluid Based Composite Medium,” Fundamentals of Heat Transfer in Electromagnetic, Electrostatic and Acoustic Fields, HTD, Vol. 248, pp. 75-82. Zhang, C., and Lloyd, J. R., 1994, “Control of Radiation Heat Transfer Through 8 Composite Window Featuring ER Fluid: A Conceptual Investigation,” Inzhenerno- F izichoskii Zhurnal, Vol. 6, No. 2, pp. 131-142. MICHIGAN STATE UNIV. LIBRRRIES 1|l1|"11111111111|11111111111WWI11111111111111." 31293016822524