. .. . x. ' . ~ ~ . . . 1 “:5: ”fix ‘ J," ’— ‘ , ‘ . V - J. ‘. ‘ , . , '31:}? : “ WM f ”(Eff-'2? 1 t "tide-‘5 «1&4'5": ‘In Ln fl‘tp‘ -- ..1 46391. H6. 'wifirffi'fim : r“; “nan: . .g. -n v. W «(4' w". "W C .2», I ‘> THEM This is to certify that the thesis entitled The Mechanical Behavior of CVD Diamond Thin Film Diaphragms presented by William Harrison Glime III‘ has been accepted towards fulfillment ~ of the requirements for Master's degree in Materials Science m Major professor Date SCH“ 12/ Hqfi/ 0-7639 MS U is an Affirmative Action/Equal Opportunity Institution m ._.——-. __‘_._»~ 4 4- Illlllllllllllllll LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. OID FINES return on or before date due. DATE DUE DATE DUE DATE DUE THE MECHANICAL BEHAVIOR OF CVD DIAMOND THIN FILM DIAPHRAGMS By William Harrison Glime III A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Materials Science and Mechanics 1993 ABSTRACT THE MECHANICAL BEHAVIOR OF CVD DIAMOND THIN FILM DIAPHRAGMS By William Harrison Glime III The behavior of centrally loaded CVD polycrystalline circular diaphragms grown on single crystal silicon substrates was investigated. Circular diamond diaphragms were prepared by mechanically grinding and chemically etching the silicon substrate. The diamond diaphragms buckled due to a compressive residual stress in the diamond film. The morphology of the buckled diamond diaphragms was analyzed using laser scanning confocal microscopy (LSCM). A nanoindenter was used to apply a central point load to the diamond diaphragms and measure the resulting displacement. The load deflection behavior of the diaphragms was compared to established theoretical nonlinear plate solutions. ACKNOWLEDGMENTS I would like to thank my family for their support and encouragement as well as Dr. Eldon Case, whose performance as an advisor was exemplary. This work was funded by the State of Michigan Research Excellence Fund. iii TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES I. INTRODUCTION 1.0 Potential of diamond films - applications 1.1 History of diamond synthesis 1.2 Development of the modern CVD method 1.3 Modern CVD methods 1.4 CVD diamond growth A 1.4a The role of hydrogen 1.4b Effect of oxygen and other gasses 1.4c Diamond nucleation on non-diamond substrates 1.4d CVD diamond growth on non-diamond substrates 1.4e Diamond film adhesion 1.5 Defects and impurities: effect on film properties 1.6 Summary of the CVD diamond growth process 2.0 Residual stress in CVD diamond films 2.1 Stress due to thermal expansion mismatch 2.2 Intrinsic stress in polycrystalline diamond films 3.0 Measurment of thin film mechanical properties iv ....... ....... ....... ....... ....... II. 3.1 The biaxial bulge method ....... 36 3.2 The vibrating membrane method (mechanical resonance) ....... 40 3.3 Microbeam deflection ....... 41 3.4 Ultralow-load indentation (nanoindentation) ....... 42 3.5 Point load deflection of circular diaphragms ....... 46 4.0 Deflection theory for point loaded clamped circular plates ....... 48 EXPERIMENTAL PROCEDURE ....... 52 5.0 Diamond growth parameters ....... 52 5.1 Specimen preparation ....... 52 5.2 Pre-dimple specimen mounting ....... 52 5.3 Polishing ....... 54 5.4 Dimpling ‘ ....... 54 5.5 Pre—etch mounting ....... 57 5.5a Mounting specimens A—F ....... 58 5.5b Mounting specimens 6-21 ....... 59 5.6 Etching ....... 59 5.6a Etching specimens A-F ....... 59 5.6b Etching specimens 6-21 ....... 61 6.0 Diaphragm morphology analysis ....... 64 6.1 Laser scanning confocal microscope (LSCM) setup ....... 64 6.2 LSCM line scan analysis ....... 64 6.3 LSCM data analysis ....... 69 7.0 Nanoindentation of diaphragms ....... 71 7.1 Mounting for nanoindentation ....... 71 7.2 Diameter measurement and center point location ....... 71 7.3 Deflection of the diamond diaphragms ....... 75 7.4 Nanoindenter / specimen compliance ....... 79 8.0 Off-center diaphragm loading ....... 81 9.0 SEM, TEM, and X-ray analysis of diamond film ....... V 83 10.0 Point load deflection of circular glass discs ....... 87 III. RESULTS & DISCUSSION ....... 89 11.0 Visual, SEM, TEM, and X-ray diffraction observations ....... 89 12.0 Diaphragm morphology - buckling relief of stress ....... 100 13.0 Deflection behavior of point loaded circular glass discs ....... 106 14.0 Deflection behavior of point loaded diamond diaphragms ....... 113 14.1 Repeated loading of diamond diaphragms ....... 138 14.2 Off-center diaphragm loading ....... 139 15.0 Analysis of load at failure for the diamond diaphragms ....... 141 16.0 Parameters influencing diaphragm deflection behavior ....... 144 IV. CONCLUSIONS ‘ ....... 145 V. FUTURE STUDIES ' ....... 147 VI. CLOSING REMARKS ....... 149 APPENDICIES ....... 150 A. Equipment parameters ....... 150 B. Nanoindenter calibration ....... 151 C. Seeding of substrate ....... 155 D. Grain boundary relaxation calculations. ....... 156 LIST OF REFERENCES ........ 157 LIST OF TABLES Table 1. Material properties of diamond. Table 2. Summary of common diamond deposition processes [27]. Table 3. Reported elastic modulus values for CVD polycrystalline diamond films. Table 4. “Values for constants A and B used in approximate solutions to the point load deflection of circular plates (Equation 4.2). Table 5. Parameters for MPCVD diamond depositon on wafer#2 [126]. Table 6. Results of LSCM line scan analysis of diaphragm specimen A. Table 7: Radius, thickness, and load at fracture for point loaded circular diamond diaphragm specimens. Table 8: Specimen parameters and constant values for fitting of load- deflection data to Equation 14.2*. LIST of FIGURES Figure 1. Crystal structure of diamond [1]. Figure 2. Carbon phase diagram [1]. Figure 3. Schematic of a hot filament assisted chemical vapor deposition (HFACVD) apparatus [7]. Figure 4. Schematic of a bell jar microwave plasma assisted chemical vapor deposition system [37]. Figure 5. Schematic of a DC plasma jet CVD apparatus [44]. Figure 6. Schematic of a combustion flame CVD apparatus [46]. Figure 7. Simplified schematic of possible nucleation and growth kinetics for diamond CVD on a non-reacting surface [60]. Figure 8. Diagram of columnar grain growth in CVD diamond films [60]. Figure 9. Room temperature thermal strain of various materials as a function of initial temperature [4]. Figure 10. Strain in diamond film due to thermal expansion mismatch with the silicon substrate as a fimction of deposition temperature [4]. Due to the relative change in coefficients of thermal expansion for diamond and silicon, the thermally induced strain for diamond grown on silicon substrates peaks at deposition temperatures of approximately 700°C. Figure 11. Diagram of biaxial bulge apparatus used by Cardinale et al. Figure 12. Schematic diagram of vibrating membrane apparatus used in the study by Berry et al. [95,106]. Figure 13. Schematic diagram illustrating the deflection of the free standing polycrystalline diamond cantilever beam by Davidson et al. [99]. Figure 14. Loading/unloading curve obtained from ultralow-load indentation on diamond film DT—26 (Table 3), redrawn from [6]. viii Figure 15 . Relative position of specimens 6 - 21 with respect to half- wafer#2. Portions of the half-wafer which were not used in the deflection study consists of specimens which were destroyed during processing or fragments unsuitable for producing diaphragm specimens. Figure 16. Mounting of diamond/silicon specimen on dimpler base using therm0p1astic wax. Figure 17. Schematic of dimpling process. Figure 18. Diagram of dimpled region in the silicon substrate for specimens A - F. Specimens obtained from wafer#2 (6-21) have a 540 mm substrate and are dimpled to a depth of 490 mm in a similar fashion. Figure 19. Diagram of pre-etch mounting procedure for specimens A - F. Figure 20. Diagram of pre-etch mounting procedure for specimens 6 -21. Figure 21. Optical micrographs of diaphragm specimens. (A) Specimen A from wafer#1, indented into the substrate side of the diamond. (B) Specimen 16 indented into the faceted (top) diamond surface (fractured by indentation). Figure 22. Micrograph of diamond diaphragm specimen mounted on glass base and support. Figure 23. LSCM line scan profile of diaphragm specimen A. The bi- model morphology is attributed to circumferential buckling. Figure 24. Position of line scans across diamond diaphragm specimen A taken during LSCM analysis of the diaphragm morphology [122]. Figure 25. Schematic of line scan profile depicting buckled length measurement technique. Figure 26. Diagram of specimen situated in stage of nanoindenter. _ Figure 27. Schematic of diaphragm center location procedure and diaphragm radius measurement. Diaphragm is purposely drawn "non- circular” in order to illustrate the procedure. Figure 28. Schematic of diaphragm deflection (specimens A - F). For specimens 6 - 21 the substrate support was below the diamond diaphragm and the indenter tip was in contact with the faceted diamond surface. Figure 29. Schematic of off-center loading of diaphragm specimen 9. Figure 30. Diagram of “tweezer method” for the mounting of diamond film fragments for SEM examination. Figure 31. Diagram of the “hammer method” for the mounting of diamond film diaphragm fragments for SEM examination. Figure 32. Diagram of glass diaphragm fabrication process. Specimens were prepared with and without glue around the rim. Figure 33. (A) Fragment of diamond film obtained from wafer#1 and corresponding SEM surface micrograph. (B) Fragment from wafer#2 and SEM surface micrograph. (C) Fragment of diamond film obtained from a wafer used in LSCM buckling analysis [122] but not in this deflection study. Included to illustrate differences in diamond film transparency and surface morphology. Figure 34. SEM surface micrograph of diamond specimen taken from wafer#1. Figure 35. SEM surface micrograph of diamond specimen taken from wafer#2. Figure 36. SEM micrograph of specimen obtained from wafer#2 showing particles adhering to the surface of the diamond film. Figure 37. SEM micrograph of particulate. The diamond coating indicates the particles are introduced before or during the deposition process. Figure 38. SEM micrograph of back (substrate) surface of a diamond film specimen obtained from wafer#2. Figure 39. SEM micrograph of cleaved diamond film/ Si substrate obtained from specimen 13. Figure 40. SEM micrograph of fracture surface of diaphragm 16. Figure 41. SEM micrograph of silicon rim supporting diamond diaphragm A showing the tapering of the silicon substrate to the edge of the diamond diaphragm. Figure 42. TEM selected area diffraction pattern taken from a single diamond grain from a thin film specimen obtained from wafer#1. The pattern agrees with that of the theoretical <110> zone for a diamond crystal. Streaking and spot splitting indicate the presence of twins and stacking faults in the diamond grain. Spots resulting from double diffraction are also present. Figure 43. X-ray 29 scan of diamond film on silicon substrate obtained from wafer#1. Figure 44. Load-deflection behavior of Simply-supported circular glass coverslip prepared by placing the coverslip (no glue) over a circular hole in a glass plate. The compliance curve obtained from indentation into the back of the glass plate(s) was subtracted from the data obtained for the clamped and simply supported plates. Figure 45. Load-deflection plots obtained from point loading of circular glass coverslips clamped by applying superglue around the rim. Figure 46. (A) Load-deflection plot for simply supported glass beam used in the nanoindenter calibration..( B) Schematic of glass beam setup. Figure 47. Load-deflection plots for clamped glass coverslips with load normalized for radius and thickness. Figure 48. Plot of combined compliance curves taken from various diamond specimens and specially prepared compliance specimens. The compliance behavior was described by Equation 14.1 which was used to subtract the compliance contribution from the diaphragm deflection data. Figure 49. Load-deflection plot for diamond diaphragm specimen A. Figure 50. Load-deflection plot for diamond diaphragm specimen B. Figure 51. Load-deflection plot for diamond diaphragm specimen C. Figure 52. Load-deflection plot for diamond diaphragm specimen D. Figure 53. Load-deflection plot for diamond diaphragm specimen E. Figure 54. Load-deflection plot for diamond diaphragm specimen F. Figure 55. Load-deflection plot for diamond diaphragm specimen 6. Deflection to 9pm repeated three times. Figure 56. Load-deflection plot for diamond diaphragm specimen 8. Deflection to 9pm repeated three times. Figure 57. Load-deflection plot for diamond diaphragm specimen 9. Deflection to 9pm repeated two times. Figure 58. Load-deflection plot for diamond diaphragm specimen 11. Deflection to 9pm repeated three times. Figure 59. Load-deflection plot for diamond diaphragm specimen 12. Deflection to 911m repeated three times. Figure 60. Load-deflection plot for diamond diaphragm specimen 13. Deflection to 9pm repeated three times. Figure 61. Load-deflection plot for diamond diaphragm specimen 14. Deflection to 9pm repeated three times. Figure 62. Load-deflection plot for diamond diaphragm specimen 16. Deflection to 9pm repeated four times. Figure 63. Load-deflection plot for diamond diaphragm specimen 17. Deflection to 9pm repeated three times. Figure 64. Load-deflection plot for diamond diaphragm specimen 19. Deflection to 9um repeated three times. Figure 65. Load-deflection plot for diamond diaphragm specimen 21. Deflection to 9um repeated two times. Figure 66. Plot of C1 (Equation 14.3) versus t‘t/r2 for all specimens. Dashed line represents forced fit through zero, solid line represents floatiing intercept. Figure 67. Load-deflection plots for off-center point loading of diamond diaphragm specimen 9. Figure 68. Plot of load at diaphragm failure versus t2. I. INTRODUCTION 1.0 Potential of diamond films - applications. The material properties of diamond (Table 1) are very impressive but the lack of an abundant natural supply and the inability of diamond to be effectively manufactured have limited its commercial use to date. Industry has utilized the hardness of diamond as abrasives and tool coatings, but diamond’s many other properties have yet to be exploited to the same degree in practical applications. The extreme hardness, low coefficient of friction, and corrosion resistance of diamond make it desirable as a protective coating. Diamond’s unmatched thermal conductivity makes it the optimal material for heat transfer. Diamond’s wide band optical transparency makes it useful in window applications. The effective production of semi-conducting diamond would allow for active diamond electronic devices which are faster, more durable, and capable of handling more power than those using current semiconducting materials [7 -9]. The potential of diamond as an engineering material is remarkable but limited by the current methods of diamond synthesis [7,8]. The rapid advancement of synthetic diamond production over the past decade makes the future of diamond as a commercially competitive material look very promising. Diamond films are already being used commercially as speaker diaphragms in high performance audio tweeters and as windows for X-ray lithography 2 Table 1. Material properties of diamond# Property Reported value ref. Density 3.515 g/cm3 [1] Hardness* (knoop) 75 - 113 GPa [2] Elastic modulus (aggregate)* 1050 GPa [2] Poissons ratio (aggregate) 0.07 [2] Melting point (carbon triple point) ~ 4200 K [1] Thermal conductivity* 20 W/cm K [2] Electrical resistivity 1016 Q-cm [1] Band gap 5.48 eV [2] Breakdown voltage >107 V/cm [1] Dielectric constant 5.5 [1] Coefficient of friction (in air) ‘ 0.05 [1] Fracture toughness (Kc - CVD Polycrystal) 4.3 MPaVfii [3] Coefficient of thermal expansion 1 x 10'6 K‘1 [4] Electron velocity 2.7 x 107 cm/sec [1] Spectral transmission range* .225 - 25pm [1] High resistance to chemical attack (moderate temperatures) [1] High resistance to radiation damage [5] Resistant to thermal shock [6] l' 'Iype IIa natural diamond single crystal at room temperature unless noted otherwise * Designates highest of known solids Diamond crystal “types” [1,7]: Type Ia - nitrogen impurity (~0.1%) in aggregate or platelet form. Type Ib - nitrogen impurity in substitutional form. Type IIa - no (trace) impurities. Type IIb - boron impurity 3 [7,8,10,11]. Doped semiconducting diamond has been produced using Chemical Vapor Deposition (CVD) processes [7,8,12-14] and functional diamond p-n diodes have also been fabricated [12,14]. In fact, researchers have predicted that the market for synthetic diamond films will be in excess of a billion dollars by the year 2000 [7,10,11] and even that may be an underestimate given the rapid development and improvement 'of diamond synthesis techniques. The advancement of diamond technology in areas of industrial tooling would also make the manufacture of other ceramics and ' composite materials more cost effective, which is one of the major factors keeping advanced materials out of many modern markets [15]. 1.1 History of diamond synthesis In 1797 Smithson Tennant discovered diamond is a form of carbon [16,17]. Since that time scientists have been striving to produce synthetic diamonds. The history of synthetic diamond production is somewhat comical, full of fraudulent claims, misguided theories, and extravagant techniques and equipment [16,17]. The majority of early synthetic diamond research was based on the transformation of graphite into diamond under thermodynamically stable conditions, i.e. very high pressure [16]. In 1954, after more than a century of failed attempts, it was announced that synthetic diamonds had been produced from graphite by researchers at General Electric [17]. A lesser known, but perhaps more significant, approach to the synthesis of diamond was the growth of diamond under “metastable conditions” at subatmospheric pressures [18-21]. In 1953, W. G. Eversole succeeded in producing new diamond on diamond seed crystals at subatmospheric pressure using a CVD method [22 as cited in 18,23]. Successful low pressure synthesis of diamond preceded the high pressure approach but was virtually ignored at the time due to extremely slow growth rates, the eventual nucleation of graphite, and the need for a diamond substrate [18,19,21,23-25]. Since the success of Eversole, a'variety of CVD methods for the low pressure production of diamond films have emerged. 1.2 Development of the modern CVD method Graphite, not diamond, is the thermodynamically stable (Figures 1 & 2) and kinetically dominant phase of carbon at other than ultra high pressures and temperatures [7,18,19,21,23]. From the thermodynamic standpoint, the growth of graphite is favored over diamond growth under CVD conditions. The information available to early researchers of diamond growth included [18,24] 1. The spontaneous transformation of diamond into graphite is not significant at temperatures below 1300°C. 2. Carbon atoms have a high mobility on diamond surfaces at temperatures around 1000°C. 3. Homogenous nucleation of graphite from gas-phase hydrocarbons is unlikely. Early researchers of low pressure diamond synthesis predicted that carbon Figure 1. Crystal structure of diamond [1]. Pressure (GPa) 30 25 M O H 01 H O 01 _ DIAMOND High-pressure / high temperature diamond synthesis Catalytic high-pressure / high—temperature diamond synthesis 833°” GRAPHITE 2000 3000 Temperature (Kelvin) Figure 2. Carbon phase diagram [1]. 6 at temperatures of approximately 1000°C, surface carbon atoms would be more likely to migrate to and attach themselves to a diamond site rather than form a graphitic nucleus [24]. Early experiments however, produced only minute amounts of new diamond on diamond seed crystals before the nucleation and growth of graphite overwhelmed the diamond growth process [18,19]. In 1967, researchers at Case Western Reserve University, led by John Angus, showed that hydrogen effectively etches the graphitic carbon that eventually formed during the deposition process. The removal of graphite occurs without removing significant amounts of diamond [24]. Subsequent deposition and etch cycles produced substantial amounts of new diamond on the diamond crystals but the procedure was far to slow and inefficient to be of practical use. In 1977, Russian scientists Derjaguin and Fedoseev published a previously unimaginable diamond growth rate of lum per hour [26]. The Russian method combined the diamond growth and hydrogen etching processes by supplying additional energy to the gas phase via hot filaments or electric discharge, which created a greater dissociation of the CVD gasses (CH4/HZ) into radical species, thus producing enough atomic hydrogen to actively etch the graphitic carbon [8,18-20,23,26,28,29]. The process developed by Derjaguin and Fedoseev forms the basis of the diamond CVD techniques used today. 1.3 Modern CVD methods One of the most widely used low pressure diamond growth techniques is the hot filament assisted chemical vapor deposition (HFACVD) process [7,18,19,21,29-30]. The HFACVD process (Figure 3) uses a tungsten or other filament heated to approximately 2000°C to dissociate the introductory CVD gasses [7,18,19,21,29,30]. The desired substrate is placed less than 1cm from the hot filaments and substrate temperature is generally maintained between 800 and 1000 0C [7,18-20,29,31]. Uneven heating of the substrate surface by the hot filaments is a major problem for the HFACVD technique. Moving the substrate away from the filaments and giving it a positive bias with respect to the filaments provides an electron current to the substrate [18,19,21,29]. Electron bombardment enhances the removal of graphite and assists in the breaking up of gaseous-carbon species near the substrate surface [18,19,21]. Moving the substrate further from the hot filaments reduces the problem of nonuniform substrate surface heating [18,19]. Chamber pressures of 10 to 100 Torr, hydrogen concentrations of approximately 99%, and diamond growth rates of 1 to 10um per hour are typical for the diamond HFACVD process [18-20,29,31]. One drawback to the process is the introduction of impurities into the diamond film due to the decomposition of the filaments [29,32,33]. Another approach to diamond growth is plasma assisted chemical vapor deposition (PACVD). Microwave, arc and glow discharge, as well as radio frequency (RF) power are among the methods used to ‘ dissociated gas substrate hot filament -\\ .\ \ “n“? .W‘ \\ ............................................................................................... ........................................................................................... .................................................................................................. electrical connection Figure 3. Schematic of a hot filament assisted chemical vapor deposition (HFACVD) apparatus [7]. antenna waveguide fused silica bell jar ..... p ; ‘ plasma ball microwave cavity substrate copger substrate ho er and succeptor graphite with cooling system substrat- holder 5:215:31 .: . :3 '-".‘ : 2| g; g . thermocouple .5. FEW-2° 'Cvm‘. \f A‘IL ‘ gas mixture gas mixture '3? -—> vacuum pump "'S'X-I'K'iill-24'3“??? 302+. . water U ... temperature -"°’°”’°- -" 4"" control meter water Figure 4. Schematic of a bell jar microwave plasma assisted chemical vapor deposition system [37]. generate plasmas for diamond growth but the use of 2.45 GHz microwaves is by far the most common (Figure 4) [18,19,34-37]. The microwave plasma assisted (MPCVD) reaction chamber is typically a fusedosilica tube inside which the plasma of hydrogen and hydrocarbon (or other carbon containing gaseous species) is maintained [18,19,35,38]. The substrate is generally placed inside the plasma although diamond growth on substrates placed outside the plasma has been reported [18,19,27]. Substrate temperature is controlled by the relative position inside the plasma and the microwave (or RF) power output [18,19,27]. Gas composition, chamber pressures and growth rates are similar to those of the HFACVD method [18,19,27]. Magnetic fields have been used to enhance the (MPACVD) process by creating a condition of electron cyclotron resonance (ECR) [39-41]. The ECR condition provides higher plasma densities and electron energies than ordinary microwave plasmas creating a larger and more uniform plasma which can deposit diamond over a greater area [41]. Arc discharge plasma jet CVD (ADPJCVD) processes have been used to grow polycrystalline films at rates in excess of 900nm per hour [27]. The ADPJCVD process (Figure 5) involves the dissociation of introductory gasses using an arc discharge [18,42-44]. The are discharge plasma has an approximate temperature of 5000°C which readily dissociates the CVD gasses [43]. The process is limited to a relatively small deposition area (~10mm) and film non-uniformity as well as poor temperature control due to the high gas temperatures w. 10 pose significant problems for this process [18,27]. High quality diamond has been synthesized on various substrates using a combustion flame method (Figure 6) [27 ,45,46]. Using an oxy-acetylene torch with a reducing (excess hydrocarbon) flame provides an atmosphere containing atomic H as well as hydrocarbon and radical species that is similar to environments in other CVD methods [21,45,46]. A water cooled substrate holder controls the temperature of the substrate which is heated by the flame. Diamond growth using this method has been reported with substrate temperatures as low as 370°C [45]. Typical growth rates in excess of 30pm per hour and the relative simplicity of the technique make the combustion method very appealing [45,46]. Like the arc discharge plasma jet method, the combustion flame process is plagued by poor substrate temperature control and non-uniform film thickness [18]. A high degree of non-diamond carbon is generally present in diamond films grown using combustion flames [45]. A variety of CVD methods have successfully synthesized diamond from the vapor phase. Table 2 summarizes growth parameters for various diamond CVD processes. The continuing improvement of these methods will undoubtedly increase the marketing potential of synthetic diamond films. 1.4 CVD diamond growth Modern diamond CVD techniques use mixtures of carbon containing Species (typically CH4 but gasses such as acetylene and 11 H2 + CH4 DC power supply cathode / anode 4-—— plasma jet substrate \ “/9 §\\\ [ ill cooling water substrate holder Figure 5. Schematic of a DC plasma jet CVD apparatus [44]. Oxygen —‘> . acetylene —> -; ijf-i‘y‘ inner core outer flame / acetylene feather substrate \ acetylene feather ;. .-.-.-.'.-.-.-.-.'. """"""" .'.-.~‘.'r7:".-7H.'5?."..-.'.-‘.-T’.".'T".‘7".‘7'7-.‘T’I‘TV.‘".CCI ..... '. . . .-.'.- ...... -.'.-.'.'.‘.‘ -.'.;. water if: i; 1:: water cooled substrate holder " 'n'. 3‘3 outer flame Figure 6. Schematic of a combustion flame CVD apparatus [46]. 12 Table 2. Summary of common diamond deposition processes [27]. Results growth deposition rate area advantages drawbacks Method (um/hour) (cm2)* heated 0 3 2 100 Simple contaminations filament ' ' large area stability, no 02 microwave 1 - 3o 40 quality, Stablhty rate, area (reasonable rate & area) 1 . highest rate, area, stability DC p asma Jet 930 < 2 quality homogeneity combustion . , , flame 30 - 100 < 1* ample area, stability low pressure simple . DC discharge < 0-1 70 large area qualltyi rate arcs, * Larger deposition areas are possible if substrate is scanned or mulitple flames, etc. are used. ‘ __._ .- ._ ...._ ___...__._..._.4~.... . - . _. .— 13 COx are also used) and H2 as the introductory gas [27 ,38,43]. Upon entering the deposition chamber, the introductory gas undergoes significant dissociation into atomic hydrogen, acetylene (CH2), methyl radicals (CH3), and a variety of other hydrocarbons and radical species [47]. There has been considerable debate over which carbon species is (are) responsible for CVD diamond growth [31,38,48-56]. The various low pressure CVD techniques (i.e. hot filament, plasma assisted, flame combustion, etc.) produce similar polycrystalline diamond deposits, indicating a single growth mechanism which is independent of the method used [21,27,53]. Acetylene, the methyl radical, methane, ethane, and other carbon species have been proposed as the depositing hydrocarbon [47,50,52,54,57]. Research has indicated, however, that acetylene and the methyl radical are the most likely diamond forming species [31,56,57]. Researchers at Case Western Reserve University use carbon-13 isotopic labeling to discern the growth species in a typical hot-filament CVD process [50]. The CWRU study used an HFCVD process in which hydrogen gas wasintroduced into the reaction chamber above the filaments. 13CH4 and 12C2H2 were introduced from a tube directly above the substrate in order to reduce the amount of dissociation and mixing of the isotopically labeled species. The amount of C13 in the deposited film and in the gases (CH4, CH3, and C2H2) directly above the substrate was measured. (111), (100), and (110) single crystal diamond as well as polycrystalline diamond were used as substrates. The C13 content of the deposited film grown on every substrate agreed with the measured C13 content in ._.-——___——_n— 14 the CH3 directly above the substrates. The amount of C13 measured for the acetylene above the substrate was significantly lower than that of the deposited diamond on each substrate. This study provided experimental evidence that the methyl radical is the primary growth species during diamond CVD on the three {(111), (100), and (110)} primary diamond faces. Several other studies have also supported the methyl radical (CH3) as the primary diamond forming species [49,55,56,58]. 1.43 The role of hydrogen Before the incorporation of hydrogen into the process, the growth of useful diamond films by CVD was unattainable. Three major mechanisms have been proposed describing the role of atomic hydrogen in diamond CVD processes. First is the “enhanced” etching of graphite relative to diamond by atomic hydrogen [23,29,47,59]. The early work by Angus showed the etching rate of graphite to be up to 500 times higher than that of diamond in an atmosphere of atomic hydrogen [24]. Even though graphite can readily form during diamond CVD, graphite can be even more readily etched by atomic hydrogen, leaving diamond to grow virtually undisturbed. A second way in which hydrogen promotes the growth of diamond over that of graphite is by maintaining the dangling bonds on the diamond surface [23,29,47]. Without hydrogenated bonds, the high energy of the diamond surface would promote reconstruction into a ‘ 15 graphitic carbon layer [23,29,47]. Surface stabilization reduces the critical size of diamond nuclei by reducing surface energy [60]. Atomic H also breaks up the larger gas phase hydrocarbons (i.e. aromatic rings) which are believed to be the “precursors of carbonaceous solid phases” [47]. The dissociation of these hydrocarbons into smaller, diamond forming species promotes diamond growth while inhibiting the formation of graphite or other non-diamond carbon phases. Dissociation of H2 into atomic H is a major function of the hot filaments, microwave plasmas, DC arcs, etc. found in modern diamond CVD processes. Increased amounts of atomic H produced in the hot filament or plasma region create a “supersaturation” of H above the substrate [53]. The increased population of atomic H allows for the aforementioned roles of hydrogen in the CVD process. The amount of atomic H produced by the different systems can vary, but its presence in adequate quantities is one of the most critical factors in successful diamond production [18,23,47,53,60,61]. 1.4b Effect of oxygen and other gasses The addition of oxygen in small quantities (< 5%) to the diamond CVD introductory gas has been shown to improve growth rates, extend the temperature region of diamond formation, and produce diamond films of higher quality than those grown without the presence of 02 [37-39,48,58,59,62,63]. A recent study [49] has provided experimental evidence that oxygen enhances the diamond 16 growth process by producing an increased amount of atomic hydrogen and CH3 in the CVD gas. Assuming CH3 is the diamond forming species, a greater concentration would increase the diamond growth rate. An increase in the concentration of atomic H promotes the removal of non-diamond carbon species. Oxygen has also been shown to provide selective etching of non-diamond carbon phases [36]. A study by Kowato and Kondo [58] indicated that oxygen suppresses the formation of acetylene, which their experimental evidence suggests is a precursor to graphitic or amorphous carbon. The increase in growth rate and improved quality of the films in the study was attributed to the reduced formation of non-diamond carbons (not just the enhanced preferential etching) as well as an increase in the population of CH3. The addition of water vapor (H20) has been shown to improve the diamond growth rate [64]. The enhanced growth rate is attributed to the same mechanisms as the addition of oxygen due to the dissociation of the water vapor into H and OH radicals. Doping with boron has also been shown to increase the crystalline quality of CVD diamond films [9]. “Enhanced surface diffusion of carbon atoms” on B doped diamond has been suggested as the mechanism creating an improvement in the quality of the CVD diamond by the addition of boron [9]. Significant quantities of nitrogen can be incorporated into the diamond lattice but due to its negative effect on the thermal, optical, and electrical properties of diamond, nitrogen is generally an 17 undesirable impurity [1]. Recent experimental evidence has suggested that nitrogen can promote carbon sp3 (diamond) bonding, which could aid in the preferential growth of diamond verses other forms of carbon [66]. The addition of any element that can increase the amount of atomic hydrogen, selectively etch non-diamond carbon deposits, improve surface mobility, or increase the population of the diamond forming carbon species could improve the diamond CVD process. 1.4c Diamond nucleation on non-diamond substrates CVD diamond has been successfully grown on a wide variety of substrates including metals, ceramics, glass, even graphite and “buckyballs” [20,23,27,67]. An issue that is still being widely investigated is the mechanism of diamond nucleation on non- diamond substrates. Several mechanisms of diamond nucleation on non-diamond surfaces have been proposed and experimentally observed but the nucleation and interfacial mechanisms of diamond growth are not well understood [18,19,23,60]. Abrading the surface of the substrate before diamond deposition has been shown to increase the diamond nucleation density by several orders of magnitude, which is attributed to an increase in the area number density of surface defects on the substrate [26,33,53,68- 71]. Abrasion with diamond powders is presumed to leave behind tiny diamond crystallites imbedded in the surface of the substrate [20,62,69,70]. Nanometer scale diamond crystallites or “seeds” have ..——__-—_—_._.__ ._—__ _..... 18 been observed on the substrate surface by Iijima et al. using transmission electron microscopy (TEM) and an extraction replica method [70]. TEM observation of these particles before and after a Short (5 min.) CVD diamond deposition show an increase in particle size and a transition from irregular to polygonized shapes, evidence that the diamond “seeds” provide homogenous nucleation sites for CVD diamond crystals [70]. Abrading the substrate surface using non-diamond grit has been shown to produce a high diamond nucleation density and “seeding” with diamond particles without abrasion has also produced high nucleation densities indicating both mechanisms of diamond nucleation are effective [69]. Detailed patterning of diamond growth on silicon has been achieved using photolithography and an argon ion beam to “roughen” the Si surface in a desired pattern [71]. Diamond readily nucleated on the surface treated with the ion beam while untreated surfaces remained virtually diamond free during HFCVD diamond growth [71]. Patterning of CVD diamond growth has also been accomplished by controlled “seeding” of the substrate with diamond particulates [127]. Abrasion and seeding processes consistently produce the highest diamond nucleation densities and are used in most diamond CVD processes. When carbide forming materials (i.e. Si, Ti, W, Mo, etc.) are used as substrates, the formation of a carbide interlayer appears to be a necessary step in the nucleation of diamond [18,26,69]. Nucleation densities on carbide forming substrates have been observed to be one or two orders of magnitude higher [26] than on substrates where 19 carbides do not form (i.e. Cu, Ni). Homogeneous nucleation of diamond powder in the gas phase has also been observed in diamond CVD atmospheres [61,72]. Gas phase nucleation can “seed” the substrate surface providing sites for diamond crystal growth. Researchers at Ford Motor Company [73] have proposed the existence of a new form of “metallic carbon”, which is harder but less dense than diamond, that acts as an intermediate stage in CVD diamond growth. The proposed material, named H-6 carbon, consists of threefold coordinated carbon which is topologically related to diamond and can continuously transform into diamond without breaking or crossing any bonds. The existence of H-6 carbon is supported by experimentally observed conducting surface layers on CVD deposited homoepitaxial diamond. The predicted mechanism of diamond growth through an H-6 intermediate is consistent with observed growth rates and morphologies on various diamond surfaces and may provide “a missing link” in the understanding of diamond CVD [73]. Short range heterogenous nucleation of CVD diamond on silicon [18,74-76] and B-SiC [74,7 7] has been shown to exist making the eventual heteroepitaxial growth of diamond on non-diamond substrates look promising. In a study by J eng et al. [7 6] local heteroepitaxial diamond growth on silicon substrates was achieved using an in situ surface pretreatment within the MPACVD chamber prior to diamond deposition. The surface pretreatment produced randomly distributed etch pits on the surface of a (100) silicon 20 substrate. Diamond was then grown on the pretreated surface using a conventional MPCVD process. The resulting diamond deposit showed areas of randomly oriented diamond crystals and in many locations, large (100nm) pyramid shaped cubic crystals oriented in the [110] substrate direction were observed. Local heteroepitaxy and the high quality of these large crystals was verified by X-ray diffraction, SEM and micro-Raman spectroscopy. The surface pretreatment is believed to produce oriented diamond nucleation, providing favorable conditions for heteroeptiaxial diamond growth on the silicon substrate. Adhesion, residual stress, crystal defects, and grain size are related to nucleation. A better understanding of diamond nucleation mechanisms is necessary if CVD diamond films are to reach their full potential [24,74,78,79]. 1.4d CVD diamond growth on non-diamond substrates Figure 7 illustrates a likely process for CVD diamond nucleation and growth. Once diamond nuclei have formed, surface diffusion of carbon on the substrate and gas phase addition of carbon on the diamond surface affect the growth of the diamond crystallite [18,23,60]. The individual crystallites or “islands” grow together covering the substrate surface and forming a continuous polycrystalline diamond film [60,80]. Further growth of the crystallites by the gas phase addition of carbon increases the thickness of the diamond film and results in a columnar grain DIFFUSION i CH3 CH4 removal H0 , m metas o / ° 44444 me 99999 cluster 99999 ° 4+ " 4+ 4+ IN.” A AAAAA U 0000 A .00 e... .00 .0 Figure 7. Simplified schematic 0 possible nucleation and growth kinetics for diamond CVD on a non-reacting surface [60]. O D O D ‘0 . I o diamondfilm ontm fie u -' :. x i :A . '. 8. Diagram of columnar grain growth in CVD diamond films [60]. 22 structure (Figure 8) [60,80,81]. CVD diamond growth has been shown to be most stable on (111) and (100) faces depending upon deposition conditions [18,23,74,20,81]. Diamond growth; on crystallites with preferred orientations (larger respective stable growth surface)’will be faster than on crystals with exposed surfaces less favorable for diamond growth. The enhanced growth of “preferred” crystals and the extinction of crystals whose surface or orientation is unfavorable results in an increase in the effective “grain size” as the thickness of the film increases (Figure 8). The columnar morphology and increasing grain size with thickness is commonly observed in CVD diamond films grown by the various techniques [80-82]. 1.4e Diamond film adhesion The use of diamond films as protective and wear resistant _ coatings is critically dependant on the ability of the films to adhere to the desired substrate. Reports on the adhesion of CVD diamond are somewhat limited and very little quantitative data on diamond film adhesion has been provided [79]. Substrate surface preparation (i.e. abrasion) appears to be a critical factor in improving the adherence of CVD diamond films as measured by scratch tests or similar methods [15,72,7 8,7 9]. Carbide forming materials, when used as substrates, generally provide better adhesion than non- carbide forming substrates which may be attributed to improved chemical bonding [78,7 9]. 23 Metal and diamond composite films have been produced using CVD methods [42,83]. One method combined a plasma jet CVD process to grow diamond crystals combined with an intermittent plasma Spray of metal (W, Mo, Cu, Ni, etc.) by introducing metallic powders to the plasma jet [42]. By continually decreasing the amount of powder introduced, films consisting of a diamond-metal mixture near the substrate and strictly diamond near the film surface were deposited [421]. Diamond-metal composite films were successfully produced using diamond deposition through RF plasma CVD followed by metallic electroplating to fill the space between diamond crystallites and further deposition of diamond by RFPCVD to produce a continuous diamond film [83]. The adhesion of the diamond-metal composite films to their respective substrates is improved through the reduction of thermal expansion mismatch between the diamond-metal composite layer and the metallic substrate as well as bonding between the deposited metal and the substrate [42,83]. The adhesion of the diamond crystallites produced with these methods to the substrate was in some cases improved to the point where the diamond crystals “will break rather than separate from the surface” [83]. The effect of surface abrasion, nucleation density, residual stress, chemical bonding betWeen the diamond and substrate, epitaxial matchup, and other parameters on the adhesion of the films must be studied further if diamond films are to be extensively utilized in tribological and other applications where adhesion is critical. 24 1.5 Defects and impurities: effect on film properties Defects are common in diamond films grown by the various CVD methods [84,33]. Crystal defects (i.e. twinning and dislocations), voids, microcracks, inclusions, and various impurities are commonly observed in CVD diamond films [33,72,74,75,7782,84-86,this study]. Twinning, dislocations, stacking faults, and other lattice disorder observed in CVD diamond crystals form to release stresses accumulated during growth [7 2,86]. Nucleation twins are also commonly observed in CVD diamond films [26,72,75]. Defect formation has been associated with the growing together of many “subgrains” which have different directions of crystal growth. As these diamond subgrains cross or combine, disordered regions are formed. The intrinsic stress in the diamond crystals has been partially attributed to the accommodation of misfit between these subgrains (i.e. twin boundaries) [50,75,86]. Lattice defects have been observed to increase with increasing gas phase carbon concentration [75,86]. Substrate temperature has been observed to affect defect concentrations by altering the surface mobility of carbon [18,74,86] and gas flow rate has also been shown to influence defect density which is attributed to the change in residence time of the diamond forming species [35]. A strong relationship between diamond film surface morphology and crystal defects has also been observed [18,50,67,74,85,86]. Crystal defects, mainly twinning, are predominant on {111} diamond surfaces whereas {100} surfaces are relatively defect free [18,46,50,72,75,82,86]. Growth rates on {111} 25 diamond surfaces, however, are generally observed to be higher than that of {100}, which has been attributed to the high defect density of the {111} surface promoting diamond nucleation and growth [18,50,82,87]. Impurities, either incorporated into the diamond lattice or occurring as inclusions or Second phases, have been observed in CVD diamond films [9,18,32,74,77]. Substitutional impurities distort the diamond lattice, creating residual stresses in the diamond crystal. Nitrogen and boron are the only foreign atomic species that are known with certainty to be incorporated with a significant degree into the diamond lattice [1], but hydrogen [7 7] may also be a substitutional impurity in diamond. The development of p-type semiconducting CVD diamond by doping with phosphorous [14] indicates that other atomic species may also form substitutional solid solutions with carbon in the diamond lattice. The presence and nature of hydrogen impurities in diamond is difficult to identify using current technology, but its effect on crystal and film quality is believed to be significant [32,88]. Hydrogen plays an integral part in the CVD growth of diamond and its effects as an impurity in diamond films needs to be studied further. Impurities of amorphous carbon, graphite and various carbides, for example, have been observed in CVD diamond films as inclusions. Amorphous and graphitic carbon are continually deposited during . the diamond CVD process [18,19,23,24]. Atomic hydrogen etches graphite up to 500 times more efficiently than it removes diamond [24]. Depending on deposition conditions, however, measurable 26 quantities of non-diamond carbon phases often exist in CVD diamond films. Raman spectroscopy is approximately 50 times more sensitive to sp2 (graphitic) type carbon and is often employed in analyzing the non-diamond carbon content in CVD diamond films [34,89]. Etching of the substrate and chamber material in the diamond CVD environment has been observed [18,80]. Substrate, or other foreign material entering the CVD gas can also introduce impurity phases (i.e. carbides) in the depositing diamond film [18]. The nature of the diamond crystal growth (i.e. “islands” to continuous film to columnar structure) leads to the formation of voids in the polycrystalline films [60,this study], and residual stresses in the diamond films may lead to the formation of microcracks [50,85]. ' Texturing of the crystals in CVD diamond films has been observed by many researchers [35,7 5,81,82,84,85]. Crystal orientation with <110> directions normal to the surface of the substrate is most common to polycrystalline diamond films grown by the various CVD methods [35,75,81], but <100> and higher order texturing, as well as random orientation has also been observed [85]. Texturing in CVD diamond is a function of nucleation and the preferred diamond growth planes. The mechanical properties (i.e. elastic modulus and especially hardness) are orientation dependant [1,2,85]. Strong texturing will therefore influence the mechanical behavior of CVD diamond films. It is also interesting to note that natural diamond is composed of approximately 99% C12 and 1% C13 [1]. Diamonds synthesized from 27 a highly pure (99.9%) C12 source have shown a 50% increase in thermal conductivity, and calculations indicate that ~100% C12 diamond would have a thermal conductivity near 50 W/cm K, which is over twice that of “normal diamond” [1,90]. Optical and electrical properties of diamond are also expected to improve as the C13 content is reduced [1,130]. Increasing the C13 content is expected to increase the already unparalleled hardness of diamond [11]. CVD diamond films have been grown with high C13 content (20 to 99%) [50,91] but a study of the mechanical properties on 13C diamond films has not been presented. The presence of lattice defects, impurities, voids, microcracks, and texturing will most certainly effect the electrical, thermal and mechanical behavior of CVD diamond films. An intricate relationship between deposition parameters, residual stresses, growth planes and morphologies, texturing, and the presence and population of defects exists. Understanding this relationship is critical to the development of diamond CVD processes. 1.6 Summary of the CVD diamond growth process Spear and Frenklach summarize diamond CVD experimental observations as follows [20]: i. Gas activation is absolutely required. ii. The chemical nature of the gaseous carbon precursor does not appear to be critical to CVD diamond growth. 28 iii. The various methods used produce similar diamond deposits. iv. The presence of hydrogen is required for efficient diamond growth. v. The addition of oxygen can improve diamond growth. vi. Acetylene and primarily the methyl radical (CH3) have been indicated as the predominant growth species. vii. Co-deposition and formation of graphite accompanies diamond growth. viii. The temperature dependance of diamond growth rate initially increases with temperature and then decreases. This behavior is a result of the competition between diamond and graphitic growth. ix. Diamond nucleation rate is strongly affected by substrate preparation method. x. {111} and {100} faces dominate the surface morphology of the films and twinning often occurs on {111} surfaces. 29 2.0 Residual stress in CVD diamond films Residual stresses have been shown to affect the physical properties of materials [92,93]. Residual stress in diamond films results from the combined contribution of intrinsic or deposition induced stresses and extrinsic or stresses due to thermal expansion mismatch with the substrate [94,95]. Both compressive and tensile room temperature net residual stresses have been reported for polycrystalline diamond films [89,94,95] but the majority of the diamond films reported in the literature had a residual compressive stress at room temperature. 2.1 Stress due to thermal expansion mismatch The relatively low thermal expansion of diamond [4] dictates a compressive thermally induced stress at room temperature for films grown on practical substrates under current CVD conditions. Figure 9 gives the relative thermal expansion of diamond compared to other common materials. Silicon is the most, common substrate used in the growth of CVD diamond films and was the substrate used for films in this study. To compute the thermally induced residual strain at room temperature for diamond grown on silicon, the thermal expansion polynomials [4] for diamond and silicon were compared (Figure 10). Using typical substrate temperatures for diamond CVD growth (600- 1000°C) the predicted strain induced by thermal expansion mismatch 30 °°9 _THERMAL STRAIN 0.8 - r 0'7 "’ Alumina smcon 0.4 -— 0-3 '- Diamond R. T. thermal strain (96) 0.2- 0.1 - 0.0 . .1 i 1 l Ml 1 l 1 l 1 l 1 l 1 l 1 l 1 l 1 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500 initial temperature (K) Figure 9. Room temperature thermal strain of various materials as a function of initial temperature [4]. R. T. thermal strain (96) 31 0.06 DIAMOND FILM THERMAL STRAIN 0.04 - Thermal expansion polynomials [4] Diamond: 0.02 " AL .. — = -o.010 - 5.911x10'°T + 3.320x10‘7T2 - 5.544x10'11T3 L Silicon: TA.L— = -o.071 + 1.887x10'4T + 1.394x10'7 T2 - 4.544x10'11T3 0.00 J 1 l l l l l l I l l L 1 J l L 1 1 1 l l 1 J l 250 500 750 , 1000 1250 1500 Deposition temperature (K) Figure 10. Strain in diamond film due to thermal expansion mismatch with the silicon substrate as a function of deposition temperature [4]. Due to the relative change in coefficients of thermal expansion for diamond and silicon, the thermally induced strain for diamond grown on silicon sub- strates peaks at deposition temperatures of approximately 700°C. 32 between silicon and diamond is roughly 5 x 10“. Assuming the silicon substrate has a thickness several times that of the diamond film, which is generally the case, then the majority of the expansion mismatch will be accommodated by the diamond film [96,97]. The thermally induced residual stress at room temperature can be calculated by multiplying the value for the residual strain by the biaxial elastic modulus for diamond (~1150 GPa) [94]. Due to the relative change in the thermal expansion coefficients between diamond and silicon with increasing temperature the thermal expansion mismatch at room temperature between diamond and silicon reaches a maximum at a deposition temperature of approximately 700°C (Figure 10). Increasing the substrate temperature during deposition above 7 00°C will result in a decrease in the room temperature thermally induced stress in the diamond film. In fact, based on the thermal expansion of diamond and silicon [4], growing diamond films on a silicon substrate at approximately 1350°C would result in a net thermally induced room temperature residual stress of zero. Unfortunately, a 1350 °C substrate temperature is well beyond the range of diamond CVD growth parameters [10-20,23,27,74]. 2.2 Intrinsic stress in polycrystalline diamond films Intrinsic or “deposition stress” in polycrystalline films can be the result of epitaxial mismatch, impurity, surface energy, interface energy, grain boundary energy, defect density, and growth mode 33 effects [92]. For relatively fine grained polycrystalline thin films, grain boundary density may play the major role in the development of growth stress [92,94,98]. The grain boundary relaxation model, developed by Finegan and Hoffman is commonly used to describe the intrinsic stress in polycrystalline films [92,94,98]. The grain boundary relaxation model is based on the following physical argument described by Windischman [94,98]. As the film grows from isolated nuclei, to crystallite “islands”, to a continuous film, interatomic attractive forces across the gaps and pores between contiguous grains cause elastic deformation (relaxation) of the grain walls. This relaxation is balanced by intragrain tensile forces imposed by adhesion to the substrate. By relating the intragrain strain energy to the difference in the surface energy of the adjacent crystallites and the energy of the resulting grain boundary the following approximation for the intrinsic stress was derived: r0 = distance of closest approach between atoms r0 E d = grain size 0 " [3“: (1-v)’] E = elastic modulus (2'1) v = Poissons ratio In a study by Windischman et al. [94] on the intrinsic stress in ‘ CVD diamond films, the grain boundary relaxation model accurately estimated the intrinsic stress in polycrystalline diamond films grown on silicon substrates which was experimentally determined from the radius of curvature of the silicon substrate after diamond deposition [94]. Windischmanns calculation of the stress due to thermal 34 expansion mismatch between the diamond film and silicon substrate does not agree with the thermally induced strain predicted from thermal expansion polynomials for diamond and Silicon obtained from Touloukian et al. [4]. An error in the Windischmann et al. [94] calculation of thermally induced stress at room temperature is also evident by the lack of a peak in thermal strain (Figure 10) at a deposition temperature of approximately 7 00°C. The cause of the discrepancy is not obvious from the information presented by Windischmann et al. but the error resulted in an underestimation of the thermally induced residual compressive stress in the diamond films of approximately 40% at room temperature. In this study, thermal expansion polynomials for diamond and silicon obtained from Touloukian et al. [4] were substituted into the analysis used by Windischmann et al. (Appendix D). The substitution produced results in agreement with the grain boundary relaxation predictions. Comparing the values obtained for the intrinsic stress in diamond films used in a study by Berry et al. [95] to the intrinsic stress predicted from the grain boundary relaxation model (Equation 2.1) presented by Windischmann et al. resulted in agreement. The grain size in the films used in the studies by Berry and Windischmann were very fine (~ 0.05 - 0.7 pm) and thus one would expect a larger contribution from grain boundary effects due to the relatively high grain boundary density. As the grain size increases, the grain boundary density rapidly decreases and grain boundary contributions to the intrinsic stress in polycrystalline films will therefore be diminished allowing contributions from defects and 35 other factors to become dominant. The residual stress state for typical CVD polycrystalline diamond films is very complex. The nature of low pressure diamond nucleation and growth results in nonuniform crystal structures during the early growth stages. The potential for defect formation and the presence of second phases as the film thickness increases is also presents a problem in understanding the nature of the residual stress state in polycrystalline diamond films. Net residual stress can be effectively measured but a complex stress state through the thickness of CVD diamond films is evident by the “curling and warping” of the films upon release from the substrate [99,93,this study]. Residual stress is a major problem facing the researchers of diamond films. Understanding and controlling the nature of residual stress is essential if diamond films are to be utilized in the variety of applications for which they show potential. 36 3. Measurement of thin film mechanical properties. Biaxial bulge [100,101], ultra-low load indentation [6,102-105], mechanical resonance [95,106], Brillouin scattering [107,108], beam bending [99,109,110], plate deflection [97], and many other methods have been used to experimentally determine the mechanical properties of thin films. - Jaing et al. [107 ,108] used Brillouin light scattering to measure sound velocity in thick (~400um) polycrystalline CVD diamond films with large (~40um) grains. The measurements provided a value of 1037 GPa for the elastic modulus of the CVD diamond crystallites ([110] oriented) which is in excellent agreement with the modulus for natural diamond. The Brillouin scattering technique is unable to effectively measure the macroscopic (many grain) properties of CVD diamond thin films. For polycrystalline CVD diamond thin films, measurement of the macroscopic mechanical properties has been attempted using the biaxial bulge [100], mechanical resonance [95], beam deflection [99], and ultralow-load indentation [6,105] methods with results presented in Table 3. These methods will be discussed along with a method using the point load deflection of circular diaphragms to measure residual stress an elastic properties of SiNx thin films [97]. 3.1 The biaxial bulge method The biaxial bulge method [100,101,111] uses uniform pressure to deflect a circular diaphragm of the film under study, providing 37 pressure versus deflection information. The pressure versus deflection information combined with a mathematical model for the mechanical behavior of the diaphragm provides a value for the biaxial modulus of the diamond film. Residual stress, diaphragm morphology (wrinkling or initial deflection) and the apparent stiffness of the material can greatly affect the results of a biaxial bulge experiment [111]. Cardinale et al used the biaxial bulge method to measure the mechanical properties of polycrystalline diamond films prepared by microwave plasma assisted CVD [100]. A schematic of the apparatus used in the study is depicted in Figure 11. The films were grown on Silicon substrates and had a thickness of approximately 10 um. By etching a circular hole with a radius of approximately 1 cm in the silicon substrate, a polycrystalline diamond diaphragm, held around its periphery by the silicon substrate, was created. A roughing pump was used to apply a vacuum to one side of the diaphragm while the diaphragm deflection was measured using an optical method [100]. Cardinale et al. measured diaphragm deflections of up to 100 um. The diaphragm was modeled as a spherical cap with a correction for non-spherical deformation (equations 3.1,3.2). o = biaxial stress 7%: (3.1) e = biaxial strain P = pressure r = diaphragm radius th h = diaphragm central deflection 3 = t = film thickness (3 2) :2 B = correction factor (2/3 for spherical displacements) ' 38 Non-spherical deformation of the diaphragm was observed in the study but the details of the deformation were not presented. The correction for the non-spherical deformation was determined using a diaphragm of silicon and of glass with thicknesses in excess of 100 um and a diaphragm radius of approximately 1 cm. Using the known values of elastic modulus for silicon and glass they experimentally determined a value of 2.46 for [3 (Equation 3.2), the correction factor- for non-spherical deformations. Biaxial modulus values for two diamond films were reported and presented in Table 3. The study by Cardinale et al. [100] resulted in reasonable values for the biaxial modulus of the polycrystalline CVD diamond films but their method is highly questionable. The use of the silicon and glass diaphragms, which had a plate rigidity (Equation 3.3) [113] D = plate rigidity E t3 E = elastic modulus of film (33) D = 12(1_V2) t = film thickness v = Poissons ratio approximately two orders of magnitude greater than that of the diamond film, to estimate the behavior of thin diamond diaphragms with a similar radius seems unreliable. A recent study examining the reliability of the bulge test indicated that current biaxial bulge test models are invalid for thin films in compression due to diaphragm buckling effects [111]. The diaphragms in the study by Cardinale et al. were in residual compression and the effect of residual film compression on diaphragm deflection behavior was not addressed. 39 optical probe diamond diaphragm 4— to transducer to vacuum ——> Figure 11. Diagram of biaxial bulge apparatus used by Cardinale et al. [100]. driver diamond membrane \ /A Si wafer detector first vibrational mode E........mnnmlIIIlIIIllIllll||l||||l||llllllllllllllllmm“1...... Figure 12. Schematic diagram of vibrating membrane apparatus used in the study by Berry et al. [95,106]. 40 3.2 The vibrating membrane method (mechanical resonance) Berry et al. used a mechanical resonance technique to measure the mechanical properties of circular polycrystalline CVD diamond diaphragms [95,106]. Polycrystalline diamond films in residual tension at room temperature were grown on silicon substrates. By etching away a circular region of the substrate taut diamond membranes were created. The fundamental flexural frequency of the diamond diaphragms were determined using an electrostatic excitation apparatus (Figure 12). The tensile stress in the diaphragm was determined from its relationship to the radius, density, and resonance frequency for an ideal circular diaphragm. Measurement of the residual stress was performed at several temperatures between 100 and 700 K to determine the temperature dependence on the stress in the diaphragms. The temperature dependence of the strain in the diaphragm is due to thermal expansion mismatch between the Silicon and diamond and was calculated from the known thermal expansion for silicon and diamond (Section 1.6). Comparison of the measured temperature dependence of stress with the predicted thermal strain over the given temperature range yields a value for the biaxial modulus of the polycrystalline CVD diamond films (Table 3). Residual stresses which are nonuniform through the thickness of the diaphragm, indicated by the curling of the films upon release from the substrate [92] have been observed in CVD diamond films [99,this study]. Porosity in the polycrystalline diamond films 41 resulting in a lower density than the value for natural diamond used in the study, the “un-rigid behavior” of the silicon substrate, and diaphragm rigidity (assumed zero) are a few of the parameters which may have resulted in significant error in the measured modulus for the CVD diamond films. Also, the vibrating membrane method can not be used to study diaphragms in compression, which is common for CVD diamond films. 3.3 Microbeam deflection The deflection of cantilever microbeams has been used to measure the elastic modulus of polycrystalline CVD diamond films [99]. Davidson et al. [99] produced polycrystalline diamond cantilever beams using a novel process involving photolithography, patterning, and plasma/ chemical etching. A fabricated diamond cantilever beam 3168um x 639nm x 13pm was mounted under an optical microscope and manually loaded using a hook and weight (Figure 13). The deflection of the beam was measured with a “point to point focus technique” using the calibrated stage of the optical microscope accurate to approximately ilum. Three loads (0.2, 0.5, and 1g) were attached to the end of the beam with a hook and the corresponding deflection was measured. A deflection in excess of 300 microns was obtained for the maximum (1g) load. From the load versus deflection data and an expression for the deflection of a cantilever beam the elastic modulus of the polycrystalline diamond film was estimated. A second technique using a single load and three diamond cantilever 42 beams of various lengths was also used to estimate the elastic modulus of the films. Both methods gave values for the elastic modulus of the polycrystalline CVD diamond of 1170 GPa and 1220 GPa respectively) Which are somewhat higher than the reported modulus of single crystal natural diamond (1050 GPa). In Davidson’s study, the cantilever beams would “curl up slightly” when freed from the substrate, which may indicate non-uniform residual stress in the diamond films [93]. Warping of the diaphragms across the width of the cantilever beam would significantly increase the beam’s bending stiffness, resulting in inflated measured modulus values. How the observed curling and warping in the diamond cantilever beams affected the deflection behavior was not addressed in the Davidson et al. study. Weihs et al, used a similar method of beam deflection to measure the properties of thin film microbeams of silicon, gold, SiOZ, and other materials (diamond films were not included) [109,110]. The study utilized a nanoindenter to provide the load-deflection information. The use of a nanoindenter could greatly improve the results of the Davidson et al. or similar studies. 3.4 Ultralow-load indentation (nanoindentation) Ultralow-load indentation is often used to measure the hardness and modulus of thin films [6,102-105]. Beetz et al. [6], used a nanoindenter to measure the modulus of diamond thin films grown by hot filament CVD on silicon substrates. Several indentations 43 vvvvvvvuvvvvvvvvvvvvvuu-vvvvvuvvuuouvv-uo vvvvvvvvvvvvvvvvvvv o v 4 v v u v vvvvvvvvvvvvvvvvvvvvar-vervevavvvvvvv'vv vvvvvvvvvvv Silicon substrate Figure 13. Schematic diagram illustrating the deflection of the free standing polycrystalline diamond cantilever beam by Davidson et al. [99]. 61 LOAD (mN) co unloading 10 2'0 3'0 310 5'0 60 7'0 00 9'0 DISPLACEMENT (nm) Figure 14. Loading/unloading curve obtained from ultralow-load indentation on diamond film DT-26 (Table 3), redrawn from [6]. ,.. 44 were made across two diamond coated silicon wafers. The elastic modulus of the film was estimated using the unloading portion of the force displacement curve obtained for each indentation [Figure 14]. Averaging the elastic modulus values obtained from the individual indents produced values of 874 GPa (film with 0.5um grains, poor faceting, and a small degree of non-diamond carbon) and 536 GPa (film with ~ 611m grain size, typical diamond (111) faceting, and virtually no non-diamond carbon). Accurate calculations of modulus from ultralow- load indentation data requires “that the test surface be smooth to dimensions much smaller than the indent depth” [103], The tremendous scatter (over an order of magnitude) in the modulus values obtained from the individual indents in the Beetz study indicates the difficulty in obtaining accurate measurements on faceted diamond films using the ultralow-load indentation method. O’Hern and McHargue [105] used ultralow-load indentation to measure the hardness and modulus of CVD diamond films with surfaces smoothed uSing mechanical methods. The roughness of the diamond surface after “smoothing”, however, was not presented. The indentation response of the diamond films was compared to the indentation response of a type IIa natural diamond crystal with [100] orientation used as a reference. The measured elastic modulus for the diamond films was slightly higher (~10%) than that of the single crystal reference but a high degree of scatter in the modulus measurements from the individual indents was also observed in the O’Hern et al. study. 45 Table 3. Reported elastic modulus values for CVD polycrystalline diamond films. Study and Method ThiCkneSS Residual stress Reported elastic specimen designation" of growth of film (pm) at 298K (MPa) modulus (GPa) Biaxial Bulge [100] 864+ 38 - 9.5 - RUB-164 MPACVD 9'6 (V:.I) RDD-l97 ' 1&0 - 6.8 1059': 90 Vib. membrane [95] 1D .. 2 139 765 2D MPACVD .. 2 58 657 (V=.1) 3D ~ 2 9.4 711 Beam deflection [99] C7 13 1176 MPACVD - var. lengths 13 unavailable 1225 Nanoindentation [6] DT-26 HFACVD 1.5 - 7 . 536 unavailable DT—23 unavailable 874 *The individual specimen labels assigned by the original researchers are listed to allow the reader to directly identify each specimen in the study cited. 46 The size of the indenter tip used in ultralow-load indentation determines the volume of material that contributes to the experiment. The tip size and shape were not reported in the aforementioned diamond film indentation studies, but typical indenter tips will result in a contact area of only a few square microns. Determining the macroscopic behavior of a material, which may depend on defects and other parameters, using information gained from such a small area is questionable. 3.5 Point load deflection of circular diaphragms In a study by Hong et al. [97] the mechanical deflection of circular diaphragms of silicon nitride was used to measure the residual stress and elastic modulus of the SiNx thin films. The thickness of the diaphragms ranged from 0.09 to 0.27 pm and diameters of 1100 to 4100 um. The circular SiNx diaphragms were prepared using a photolithography and etching technique. A nanoindenter was used to deflect the circular diaphragms with a point load at their center. From the load-deflection behavior of the diaphragms in residual tension a value for the elastic modulus and the residual stress in the films was obtained using classical plate deflection theory (Section 4) with modifications for residual tension. The study by Hong et al. was limited to films in tension due to the buckling of the diaphragm for films in compressive residual stress. The technique provided reasonable values for the residual stress and elastic modulus for the SiNx thin films but the deflection behavior of 47 the diaphragm is much more sensitive to the residual tensile stress in the film than to the elastic modulus. Therefore, the use of the Hong et al. technique for the measurement of the elastic modulus of thin film diaphragms in tension is “not very accurate” [97]. The use of classical plate theory to describe the behavior of plates undergoing large deflections is also questionable as will be discussed in the following section. 48 4.0 Deflection theory for point loaded clamped circular plates Classical plate theory describes the deflection behavior of a flat, fully clamped circular plate loaded at its center with the following expression [1 13,1 14]: r = radius of plate w = Prz D = plate rigidity (Equation 3.3) ( 4 1) 161tD W = central deflection . P = point load Classical plate deflection expressions, based on Kirchhoff theory, assume the stretching and the transverse shearing of the plate is negligible [115]. These assumptions hold as long as the magnitude of the central plate deflection is a fraction of the plate thickness [113,115,116]. For larger deflections, generally greater than one-half the plate thickness, deformations in the middle surface of the plate (i.e. stretching) must be considered for accurate results [113,115,116]. Plates undergoing large (greater than one-half the thickness) deflections exhibit “geometrically nonlinear behavior” which can only be accurately described using nonlinear solutions [115]. “Geometrically nonlinear” refers to a change in the load-bearing behavior of the plate due to changes in configuration during loading and does not imply material nonlinearity. Governing equations for the point load deflection of a fully clamped circular plate were developed by Von Karman in the early 1900’s and require coupled nonlinear differential equations [117,118]. Equations of this type are generally very difficult to solve 49 and an “exact” point load plate deflection solution has “defied all efforts” [118]. Several approximate solutions to the large deflection of centrally loaded circular clamped plates have, however, been presented [113,115,117-119]. . Using a strain energy method (Ritz-Galerkin method) Timoshenko and Woinowsky-Krieger developed the following expression describing the central deflection of a circular plate with a central point load [113]: .r = radius of plate ' E = {lilate lillasktlilc modulus t = p ate t 'c ess (N + A N3 ) = B [El-i] P = point load (4.2) Et N = normalized deflection (w/t) A,B = constants (Table 4) Banerjee and Datta [117,120,121] used a modified energy expression to de-couple the governing equations and develop an approximate solution to the point load deflection behavior of a circular plate. The Banerjee et al. solution is virtually identical to the solution obtained by Timoshenko and Woinowsky-Kreiger (Equation 4.2) with only a slight difference in the constants A and B for the various plate boundary conditions [Table 4]. Schmidt [118] used a perturbation method to derive approximate solutions to the large deflection problem which are in excellent agreement with the approximate solutions of Timoshenko and Bannerjee [115]. Bert and Martindale used a numerical analysis based on the energy method to study the point load deflection behavior of fully clamped circular plates [115]. The numerical results were consistent with the approximate solutions presented in the literature. 50 Table 4. Values for constants A and B used in approximate solutions to the point load deflection of circular plates (Equation 4.2). Bannerjee [117] Timoshenko [113] Boundary conditions* A B A B Edge Clamped immovable .43 .2 17 .443 .217 fifig‘zable .11 .217 .200 .217 Edge Simply- immovable 1.26 .551 1.43 .552 su orted pp fififizable .16 .551 .272 .552 * Boundary conditions are described as follows: I“ immovable moveable Clamped edge immovable - No rotation or in-plane movement at boundary. edge moveable - No rotation at boundary, unconstrained in-plane movement Simply-supported. immovable moveable edge immovable - Free to rotate, no in-plane movement at boundary. edge moveable - Free to rotate, unconstrained in—plane movement. 51 Experimental data for the very large (>1.5 t) deflection of centrally point loaded fully clamped isotropic circular plates is unavailable in the literature so the approximate solutions to this problem have not been compared to experimental results [115]. Experimentally observed deflections tend to be larger than those predicted by the theoretical approximations (i.e. equation 4.2), especially for clamped plates, which is attributed to the difficulty in obtaining “theoretical” boundary conditions (i.e. absolutely no bending, rotation, or slipping at the boundary of a fully clamped plate) [115,117]. 52 II. EXPERIMENTAL PROCEDURE 5.0 Diamond growth parameters Diamond films for this study were grown using a microwave plasma assisted chemical vapor deposition MPACVD process. Growth parameters for the diamond film from which specimens A - F were obtained (wafer 1) are unfortunately unknown. Deposition parameters for the diamond film grown on wafer 2 (specimens 6 - 21) are presented in Table 5. 5.1 Specimen preparation The diamond coated silicon wafers were cleaved into 1cm squares using a razor blade. The cleaving process involved the placement of the wafer on a paper towel with the diamond side up. A crack was initiated on an edge of the wafer using a razor blade and then propagated by applying pressure with the razor blade. The position of the 1cm square Specimens with respect to the wafer was recorded for specimens 6 - 21 as well as the position of any large fragments that were produced (Figure 15). 5.2 Pre-dimple specimen mounting The 1cm square specimens were secured using a thermoplastic wax to the stainless steel cylindrical base of a dimpling instrument 53 Table 5. Parameters for MPCVD diamond depositon on wafer #2 [126]. Single crystal (100) p-type silicon wafer r = 5 cm, seeded substrate with diamond using a photolithographic process [127]. microwave power 1820 watts at 2.45 GHz chamber pressure 40 Torr substrate temperature 840°C H2 flow rate 400 sccm CO flow rate 3 sccm CH4 flow rate 5.5 sccm deposition time 8 hours weight gain 91 mg diamond side up Figure 15 . Relative position of specimens 6 - 21 with respect to half-wafer #2. Portions of the half-wafer which were not used in the deflection study con- sists of specimens which were destroyed during processing or fragments unsuitable for producing diaphragm specimens. 54 (South Bay Technologies, Temple City, CA) used for TEM sample thinning (Figure 16). The base was heated on a hotplate set at 75°C. A small amount of thermoplastic wax (~0.1g) was placed in the center of the base and allowed to melt, forming a bead. The base was removed from the hotplate and placed in a shallow dish of water to cool. The 1cm square Specimen was pressed into the wax and centered on the base using fingers before the wax had a chance to harden. Moderate pressure (~5 N) was maintained on the specimen using fingers until the wax completely hardened. The sample was visually checked for level with respect to the base. If the specimen appeared to be out of level, the base was reheated and the specimen was again pressed into the softened wax. - 5.3 Polishing While attached to the dimpler base, the silicon surface of the specimen was polished using polishing wheels and 511m, 0.3um and 0.05pm alumina grit respectively. Polishing of the Si surface results in a more uniform etch with less pitting or other etching related defects as compared to unpolished specimens but is not critical to the specimen preparation process. 5.4 Dimpling After polishing, the base was secured to the dimpling instrument. The dimpling of the specimens was performed using a 0.5N load, diamond/ Sl waferfl. melted wax dimpler base \ _> Figure 16. Mounting of diamond/silicon specimen on dimpler base using thermoplastic wax. ' alumina slu rry , ~ wood applicator dimpling wheel \_> \ H specimen dimpler base (20 rpm) Figure 17. Schematic of dimpling process. dimple d region silicon substrate fi diamond film Figure 18. Diagram of dimpled region in the silicon substrate for specimens A - F. Specimens obtained from wafer #2 (6-21) have a 540 um substrate and are dimpled to a depth of 490 pm in a similar fashion. 56 20rpm base rotation, 90rpm wheel speed and a slurry of glycerol and 600 grit aluminum oxide mixed 5 to 1 respectively by weight (Figure 17-18). Specimens A - F were dimpled using a brass wheel 10mm in diameter and 0.8mm in thickness. Specimens 6 - 21 were dimpled using a brass wheel 15mm in diameter and 1.3mm in thickness. The wheel was secured to the dimpling axle and the base was carefully centered beneath the wheel using the stage positioning knob. The counterbalance on the dimpling arm was adjusted to provide a 0.5N load. Base and wheel rotation was initiated and the arm lowered until the wheel made contact with the specimen. The slurry was applied to thewheel using a wooden rod until the Wheel was completely coated in slurry (~0.5cc). The depth sensor on the dimpling instrument was adjusted to correspond with the top surface of the sample and the attached digital micrometer reset to zero. The depth sensor, which consisted of an electrical contact between the dimpling arm and a peg on the micrometer was adjusted to a distance of .25mm for specimens A - F with substrate thickness of .275mm and .49mm for specimens 6 - 21 with substrate thickness of .54mm. The dimpling process was automatically terminated when contact between the sensor and dimpling arm was made. The dimpling process required approximately 20 minutes for specimens A - F and 100 minutes for specimens 6 -21. The slurry was ' removed with a water rinse followed by immersion in acetone and air drying. In order to reduce the risk of damage to the diamond film during dimpling and handling, dimple depths were maintained such that roughly 50pm of silicon substrate remained between the bottom 57 of the dimple and the diamond film. The dimpling process produced a “pseudo-hemispherical” dimple with a flat circular bottom (Figure 18). The radius of the pseudo-hemispherical dimple corresponds to the diameter of the respective dimpling wheel while the diameter of the flat dimple bottom corresponded with the dimpling wheel width (0.8mm for specimens A - F and 1.3mm for specimens 6 - 21) and any error (assumed to be less than ~.5mm) in the centering of the specimen. Repeated dimpling of the specimens using the same brass wheel produced a measured decrease in the wheel diameter. The ‘ final diameter of the 15mm brass wheel after dimpling over 15 specimens was measured to be 12.8mm. The reduction of the dimpling wheel is attributed to mechanical grinding and the rate of reduction is assumed to be linearly related to the circumference of the dimpling wheel. The resulting radius of the pseudo; hemispherical dimple is also expected to decrease with successive specimens corresponding to the reduction in dimpling wheel diameter. Samples 6 - 21 were dimpled in numerical order. 5.5 Pre-etch mounting The specimen was removed from the dimpler base by reheating on the hotplate and sliding off the sample when the wax became sufficiently soft. The specimen was soaked in acetone for 30 minutes, gently rubbed with fingers, then rinsed with acetone to remove any wax which may have adhered to the diamond surface. Following the acetone rinse, the specimen was immersed in methyl alcohol and 58 allowed to air dry. 5.5a Mounting specimens A - F (Figure 19). A 3mm section was cut from a glass tube with a 7mm inner diameter and 1mm wall thickness using a low-speed diamond saw (Isomet, Buehler, Lakebluff, IL). The ends of the glass tube were ground smooth using 600 grit silicon carbide paper. A 1cm square glass plate 1.2mm in thickness, cut from a microscope slide, was glued to one end of the 3mm tube using Torr Seal (Varian Vacuume Products, Lexington, MA) epoxy. A wooden rod was used to apply a thin continuous bead of the Torr Seal to the end of the glass tube. The tube was then pressed onto the glass plate. A 50g weight was placed on the tube to maintain pressure between the tube and base while the Torr Seal was allowed to harden for 1 hour at room temperature. A thin continuous head of Torr Seal was applied to the other end of the glass tube which was then lightly pressed onto the diamond side of the specimen and centered above the dimple. A 50g weight was placed on the base and the Torr Seal was allowed to harden for 1 hour. Torr Seal was then applied around the edges of the sample and to all exposed diamond surfaces using a wooden applicator. An additional head of Torr Seal was applied around the joints between the tube,'base and sample, to ensure the etchant could not leak into the tube. N 0 Torr Seal was applied to the silicon (top) surface of the specimen and any overflow of Torr Seal from the edges was removed from the top surface with a razor blade. The Torr Sea] was allowed to cure for 24 hours at room temperature. 59 5.5b Mounting specimens 6 - 21 A long (~8cm) glass tube (7mm inner diameter, 1mm wall thickness) was attached to the silicon substrate centered above the dimple (Figure 20) using a head of Torr Seal epoxy applied in a manner similar to that for specimens A-F. The Torr Seal was allowed to cure for 24 hours at room temperature 5.6 Etching An etching solution of 10 molar KOH was prepared by placing 11g anhydrous KOH pellets in a 50ml Pyrex beaker and adding 14ml deionized water. 5.6a Etching specimens A - F The 50ml beaker was placed in the bottom of a 2000ml jacketed beaker containing 200ml distilled water. A constant temperature circulator (Fischer Scientific model 800, , which pumped distilled water through the walls of the jacketed beaker was started and set to 60°C. A watchglass was placed over the 50ml beaker to prevent water droplets. which condensed on the lid of the jacketed beaker from falling into the etchant. The temperature of the water bath inside the jacketed beaker was monitored by a mercury in glass thermometer placed through a stopper in the lid of the jacketed beaker. After the temperature of 60 dimpled specimen diamond surface glass tube bead of Torr sea] glass base r ,,,,,,,,,,,,,,,,,,,,,,, ’1; ,,,,,,,,,,,,,,,,,,,,,,,,,,, a .......... dlfill Slde VleW Si substrate Torr Seal diamond film \ glass tub/ Figure 19. Diagram of pre-etch mounting procedure for specimens A - F. bead of Torr Seal glass tube diamond side —+ dimple Figure 20. Diagram of pre-etch mounting procedure for specimens 6 -21. 61 the bath had stabilized the circulator was adjusted to bring the bath temperature to 60°C as read by the thermometer in the bath. The sample was then placed in the beaker containing the KOH etchant. The sample was etched until a diamond diaphragm of the desired size (1mm to 2mm in diameter) was exposed. The etching process was visually monitored. Etching times ranged from 4 to 8 hours, depending upon the depth of the dimple and the final size of the exposed diamond diaphragm. After removal from the etchant the sample was placed in a 100ml beaker containing deionized water at a temperature of 60°C. The beaker containing the water and sample was removed from the jacketed beaker and allowed to cool to room temperature. After soaking for at least 1 hour, the sample was immersed in acetone and allowed to air dry. 5.6b Etching specimens 6 - 21 The ~8cm glass tube with attached specimen was hung from the side of a 250m] beaker partially filled with deionized water (~200 ml) such that the open end of the glass tube was above the water line and the specimen was immersed in water. The beaker was then placed in the 60° C bath (see previous section) and the water temperature inside the beaker was allowed to stabilize. KOH etchant (~ 2ml), as prepared for specimens A-F, was placed in the glass tube attached to the specimen using a pipette. The beaker was replaced into the 60° C water bath inside the jacketed beaker. Etching times ranged from 4 to 6 hours. Several of the specimens became “unglued” from the 62 tubes during the etching process. When this occurred the specimen was re-attached to the tube and the etching process repeated. After the desired dimple diameter was attained, the beaker containing the tubes and specimens was removed from the bath. The KOH was poured from the tubes which were then flushed with deionized water. After thoroughly flushing, the tubes were filled with deionized water and the specimen was allowed to soak 1 hour. After the etching process, the specimens were twisted off the glass tube using fingers. If the bond between the glass tube and specimen could not be broken with a moderate amount of force (less than approximately .5 ft-lb) then the glass tube was cut ~3mm above the sample using a low-speed diamond saw. After cutting the specimen was rinsed in warm soapy water to remove any cutting oil, followed by a rinse and soak in methyl alcohol. After removal from the alcohol the specimen was allowed to air dry. After drying, the open end of the attached glass tube was glued to a 1cm square glass plate following the method described for specimens A - F. Specimens which were removed from the glass tube by twisting were re-mounted on a 3mm glass tube following the procedure described for specimens A - F. After the mounting procedure was complete, specimens A-F and 6 - 21 were virtually identical with the exception of substrate thickness (257 and 540 um respectively) and the exposed surface of the diaphragm. Specimens A - F were prepared such that the substrate side of the diamond film was exposed while the faceted surface of the diamond film was exposed for specimens 6 - 21 (Figure 21). 63 150nm Figure 21. Optical micrographs of diaphragm specimens. (A) Specimen A from wafer#1, indented into the substrate side of the diamond. (B) Specimen 16 indented into the faceted (top) diamond surface (fractured by indentation) 64 6.0 Diaphragm morphology analysis Visual inspection of the diamond diaphragm specimens A - F indicated the existence of a “wrinkling” and “blistering” effect. Using a fiber-optic lamp to apply light at various angles while viewing the sample under an optical microscope provided further evidence of a buckled diaphragm morphology. A Laser scanning confocal microscope (LSCM, Appendix A) was used to nondestructively investigate the morphology of several diamond diaphragms. The Diaphragm morphology was attributed to the release of compressive residual. stress in the films by the buckling of the diaphragm. 6.1 Laser scanning confocal microscope (LSCM) setup The diamond diaphragm sample (Figure 22) was placed on the stage of the LSCM and the surface was brought into focus using an eyepiece and 10x objective. A single line argon ion laser was used as the source of illumination with the LSCM set in reflecting mode. Using a 10x objective and 20x electronic enhancement the surface of the diaphragm was viewed on an attached color monitor. The dimensions of the circular diaphragm were recorded using on-screen cursors and length readout provided by the LSCM software. 6.2 LSCM line scan analysis After the diaphragm dimensions had been recorded, the approximate center of the diaphragm was determined by the intersection of perpendicular diameters. Using the package software Q) 6i 65 provided with the LSCM a line scan was taken across the center of the diaphragm. To perform a line scan, the laser on the LSCM rasters across a single line defined by the user with an on-screen cursor. As the laser rasters the specimen the motorized stage of the LSCM drops down bringing the specimen out of the focal plane of the optical objective. The motorized stage is then stepped up in 0.1um increments taking the sample through the focal plane of the optical objective. The reflected intensity of points along the scanned line is electronically recorded in conjunction with the vertical position of the motorized stage. After the line scan has been completed, a plot of the points of highest reflected intensity and their relative vertical position is presented on the monitor (Figure 23). The points of highest reflected intensity occur when the corresponding point on the diaphragm is “in focus”. The plot of the points of highest intensity provides a cross- sectional or “in plane” view of the diamond diaphragm along the scanned line. The accuracy of the method is believed to be within i 1pm based on the thickness of the diamond film and the ~5um focal length of the 10x objective lens used. The use of an objective lens with a shorter focal length (typical for larger magnifications) would reduce this error but allow only portions of the diamond diaphragm cross-section to be viewed on the monitor. Line scan information would then have to be “pasted together” in order to analyze the diaphragm morphology making the use of a larger objective impractical (i.e. using a 50x objective would require 5 line scans to 66 2mm Figure 22. Micrograph of diamond diaphragm specimen mounted on glass base and support. Figure 23. LSCM line scan profiles of diamond diaphragm specimen A. The bi-model morphology is attributed to circumferential buckling. 67‘ span the diameter of the diaphragm). If the line scan showed the Specimen was “tilted” or out of level with the stage of the LSCM then transparent plastic “Scotch” tape was used to shim the Specimen into level. The process was repeated until the specimen was as level as possible. Having the specimen level was not imperative to produce accurate results but aided in the manual data collection procedure and provided an improved appearance of the line scan data during the data acquisition process (Section 6.3). Line scans were taken at specified intervals above and below the center line scan. After the line scans spanned the entire diaphragm the specimen was manually rotated 90° (using fingers) and the process repeated (except for the tilt correction procedure). A schematic of the line scans across the diamond diaphragm is depicted in Figure 24. All line scan information was recorded electronically by the LSCM. Deflection information could be taken directly from the LSCM monitor using the on-screen cursors. Electronic enhancement of the line scan profiles by increasing the magnitude of the vertical displacement by 5x aided in the acquisition of morphology and deflection information. Prior to the manual movement of the specimen on the LSCM stage the position of several “landmarks” or distinct features of the diaphragm and silicon rim were recorded. The landmarks could then be used to reposition the diaphragm after movement or removal from the stage. 68 1 A‘Il-L ' 2 71......“ 3 III-IIIIIK 4 III-II...- 5 III-III... 1295“!“ 6 III-II...- 7 ‘IIIIIIII' 8 ‘IIIIIIIIV 9 ullllllu “.." v 10 11 12 13 14 15 16 17 18 Figure 24. Position of line scans across diamond diaphragm specimen A taken during LSCM analysis of the diaphragm morphology [122]. lllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllll \ Pixel locations Figure 25. Schematic of line scan profile depicting buckled length measurement technique. 69 6.3 LSCM data analysis All line scan information was transferred to a graphics workstation for analysis using VoxelView (Vital Images, Fairfield, Iowa) package software. The line scan information was electronically reproduced on the workstation monitor and manually analyzed by J. Mattavi [122]. Data from the line scan profiles was acquired using pixel locations on the workstation monitor. Approximately 30 data points were taken from each line scan profile. The data points were manually entered into a data file to be analyzed with a PC. The length of the buckled line scan profile was determined by summing the distance between consecutive data points (Figure 25). The length of the line scan profile was compared to the “un-buckled” cord length to provide a value for the buckling strain along that particular cord. _ buckled length 8 — cord length (6.1) Thorough LSCM analysis was performed on only two diamond diaphragms, specimen A and a specimen obtained from a CVD diamond coated Si wafer which was not used in this study. During the early “learning” stages of LSCM morphology measurement, the central deflection of several diamond diaphragms was investigated but only limited information (i.e. general trends) on the diaphragm morphology was obtained during these measurements. No morphology measurements were performed on specimens 6 - 21. 70 Table 6. Results of LSCM line scan analysis of diaphragm specimen A [122]. line # (figure 25) arc length (um) cord length (um) strain(x10°4) data points 1 867.68 867.09 6.90 20 2 1064.25 1063.77 4.51 20 3 1163.88 1163.50 3.22 20 4 1255.47 1254.92 4.34 24 5 1252.81 1252.15 5.31 28 6 1247.38 1246.61 6.19 32 7 1183.86 1182.79 9.01 29 8 975.50 975.04 4.68 26 9 782.15 781.14 12.9 21 10 814.59 814.38 2.52 21 11 1097.44 1096.92 4.78 25 12 1172.27 1171.71 4.87 31 13 1272.10 1271.43 5.28 35 14 1280.45 1279.74 5.54 37 15 1272.08 1271.43 5.13 35 16 1172.31 1171.71 5.14 31 17 1097 .43 1096.92 4.60 25 18 814.83 814.38 5.50 21 Average apparent strain induced by buckling = 5.58 x 10‘4 71 7.0 Nanoindentation of diaphragms A Dow Interfacial Testing System designed for composite fiber debond testing was used to apply a point load to the diamond diaphragm specimens. Details of components in the system are provided in Appendix A. 7.1 Mounting for nanoindentation A phenolic ring (Buehler, Lakebluff, IL), 2.5cm outer diameter, 2.2cm inner diameter and 1.9cm in height, was filled to a level approximately 4mm below the top of the ring with Epofix epoxy (Struers, Westlake, OH). Plastic masking tape was placed across the bottom of the ring and the Epoxy was mixed and poured into the ring. The ring was placed on a level surface and the epoxy was allowed to harden for 8 hours at room temperature. The hardened epoxy produced a smooth, level surface inside the ring to which the specimen was superglued. Superglue was applied to the corners of the glass specimen base which was positioned inside the phenolic ring and pressed into place using fingers. 7.2 Diameter measurement and center point location The phenolic ring/specimen was secured in the motorized stage of the nanoindenter using three set screws (Figure 26). The diaphragm, which was parallel to the nanoindenter stage, was 72 ' . . XObjective ' indenter tip specimen _ l " /—\ phenolic ring set screw .v'.MoroRlzsn:srAGE' Figure 26. Diagram of specimen situated in stage of nanoindenter. 73 manually positioned beneath the optical microscope of the nanoindenter. The vertical position of the stage was manually adjusted to bring the diaphragm into the focus of the 50x objective lens. The optical image was electronically magnified an additional 20x (total of 1000x) and displayed on a black and white monitor. The motorized stage of the nanoindenter was equipped with manual motor controls and could also be controlled using the attached PC and keyboard. The motorized stage has an X and Y (the plane of the diaphragm) step resolution of lum and Z (vertical) step resolution of 0.04pm. The position of the stage was displayed by the stage motor controls. Using a marker on the monitor, the relative location of any point on the diaphragm could be determined from the position of the motorized stage. To measure the diaphragm diameter, the center line of the diaphragm was visually approximated. The distance across the approximate center line (diameter 1) was determined from the position of the points on the rim of the diaphragm. The midpoint of the approximate center line was determined and the diameter of the diaphragm perpendicular to and passing through the midpoint of the center line was measured (diameter 2). The midpoint of diameter 2 was determined and the diameter of the diaphragm perpendicular to and passing through the midpoint of diameter 2 was measured (diameter 3). This procedure was repeated until the center points of consecutive diameter measurements were in close proximity (20pm or less), which was usually after the second or third diameter measurement (Figure 27). Measurement of the diaphragm diameters 74 45° from the center line and passing through the approximated midpoint of the diaphragm was also performed. The average diaphragm diameter was determined from the average of the four diameter measurements. The center point of the diameter was approximated from the centers of the four measured diameters. A measurement of the vertical displacement of points on the diaphragm could be made using the focal plane of the 50x objective in combination with the motorized stage by recording the Z position of the stage at the point where the diaphragm was in focus. The focal length of the 50x objective is ~2um. Combining the focal length with human error in determining when the surface is “in focus”, limits the measurement of the initial (residual stress induced) central diaphragm deflection on the nanoindenter to approximately 4pm (100 vertical steps). The relative central deflection of the diaphragm was determined from the approximate vertical position of points around the rim of the diaphragm and the approximate vertical position of the center point. The degree of “tilt” of the specimen was also approximated from the relative vertical position of the points around the rim of the diaphragm. For specimens 6 - 21, which had the faceted (top) diamond surface exposed for indentation, the boundaries of the diaphragm could not be distinguished by the optical microscope on the nanoindenter. By turning off the light in the microscope as well as the roomlights and using a penlight to “backlight” the relatively transparent diamond diaphragm, the boundaries of the diaphragm could be distinguished by the intensity of transmitted light (the opaque silicon substrate 75 effectively blocked the light, defining the rim of the diaphragm). 7.3 Deflection of the diamond diaphragms (Figure 28) After the center of the diaphragm was located the stage was moved to position the diaphragm center beneath an indenter tip connected to the 50x objective. The position of the indenter tip with respect to the focal point of the objective lens (marker on the monitor screen) was determined by a prior calibration procedure. The Interfacial Testing System (ITS3) software package, designed for automated fiber debond testing, was used to automatically control the motorized stage position and collect the load-deflection data. The following parameters defined the motion of the stage during the deflection of the diaphragm: step size (vertical) = 0.04pm step rate = 4 steps/sec maximum allowable load = 2g loading increment = 1g (defines number of loading cycles) load drop for debond (fracture) 0.2g During positioning, the stage of the nanoindenter was moved at a rate of approximately 500 steps per second for all directions (X, Y, and Z). The stage would position the indenter tip above the center of the diaphragm and then drop down a predetermined distance (determined by the prior calibration) before attempting data 76 Si SUBSTRATE initial guess for Centerline \ djaph ragm center a proxrmated from c uster of diameter centers. Figure 27. Schematic of diaphragm center location procedure and diaphragm radius measurement. Diaphragm is purposely drawn “non-circular” in order to illustrate the procedure. substrate indenter tip /' diamond diaphragm fl stage motion faceted (top) side of diamond film Figure 28. Schematic of diaphragm deflection (specimens A - F). For specimens 6 - 21 the substrate support was below the diamond diaphragm and the indenter tip was in contact with the faceted diamond surface. 77 acquisition. The software was capable of handling only 300 steps (12pm) of vertical stage travel during data acquisition. Deflections in excess of 12pm required manual data collection or additional runs with the diaphragm partially deflected before initiating data acquisition. Contact with the diaphragm was defined by the first non-zero load measured by an electronic balance attached to the stage of the nanoindenter. During automatic measurements all load- deflection data was recorded by the computer, printed, and stored on a floppy disk. The motorized stage of the nanoindenter had a maximum vertical travel length of 40,000 steps (1.6mm). The accompanying ITS3 software allowed for only 20,000 steps (0.8mm) of vertical stage travel during positioning and data acquisition. The allowable working distance was defined by the maximum stage travel of 1.6mm. The allowed working distance for automatic data acquisition (0.8mm) was defined by the ITSS software. Included in the working distance was an “indenter clearance” providing a margin of safety between the indenter tip and the surface of the sample during the positioning of the stage. Collisions with the specimen during stage positioning could damage the indenter tip and/or motorized stage and care was taken to avoid such a collision. If sufficient clearance between the indenter tip and specimen could not be attained using the working distance defined by the ITS3 software then manual stage positioning and data acquisition was required. Manual operation of the nanoindenter during positioning and data acquisition was performed using the motorized stage controls 78 (buttons). The diaphragm measurement and centering process is as previously described. After locating the center of the diaphragm, the stage was dropped to provide clearance for the indenter tip. The stage was then moved a predetermined amount to position the indenter tip above the center of the diaphragm. The stage was moved vertically by an amount determined by the prior calibration and the additional “clearance” distance, bringing the indenter tip close to the surface of the diaphragm. The stage was then moved vertically at approximately 4 steps per second (0.04pm each) by manually pressing the vertical control key at the specified rate. Contact with the diaphragm was determined from the stage position at the first registered load. Load values were read and indicated vocally by the operator and recorded using an audio tape. The recording was done at twice normal tape speed. Replaying the tape at normal speed aided in the collection of data. By counting the number of audible “clicks” emitted by the stage control, the position of the stage at the point where the appropriate load was vocally recorded could be determined. Since the step rate is fairly slow and the step size is small the error due to human response time is negligible. The collection of data for diaphragm deflections in excess of 12pm was also performed by partially deflecting the diaphragm prior to automatic data acquisition. The partial deflection of the diaphragm was attained by increasing the vertical component of the “calibration” relationship between the focal point and the indenter tip which essentially “tricks” the nanoindenter into beginning data 79 collection from an already deflected diaphragm. The amount of “pre data” deflection induced by the indenter tip was approximately 811m (200 steps) which allowed a total diaphragm deflection of roughly 20pm. A 20pm deflection was beyond the failure point of all the diamond diaphragms tested. The data for the pre-deflected diaphragm was shifted to correspond with the data in which no initial deflection was induced before data collection. The measured ' deflection values for pre-deflection data was Shifted by an amount that brought the measured load values for the data sets in agreement for deflection values where the data overlapped. 7.4 Nanoindenter / specimen compliance The compliance of the nanoindenter and sample during the diaphragm deflection measurements was estimated using two specifically prepared specimens. The compliance specimen was prepared with the method used in preparing the diamond diaphragm specimens with the exception of the grinding and etching procedure. The silicon surface was polished using polishing wheels and alumina grit (5, 0.3, and 0.05pm respectively). The approximate center of the specimen, as determined by the underlying circular glass tube, was located and indented with a force-of 0.25N (approximately 10 times the typical load for diaphragm fracture). Silicon has been shown to exhibit anomalous behavior during indentation due to a stress induced phase transformation [123]. The anomalous behavior during indentation of the silicon substrate was also observed in this study. 80 The initial 0.25N indent provides a site where further indentations at lower loads would not be expected to cause further phase transformation. Several indents were made in the initial depression using parameters (step rate, etc.) identical to those for diaphragm deflection. The load versus deflection of the sample was recorded for each indent. The entire procedure was repeated at two additional locations in the vicinity of the sample center and on the second compliance sample. An estimate of the elastic response of the silicon in the vicinity of the indenter tip during indentation of the compliance specimen was made by comparison with low-load indentation data obtained by Pharr [123]. The elastic response of the silicon substrate during indentation in the Pharr study was less than 0.1um for the range of loads used in this study (less that 40 mN). Although the indenter used in the Pharr study (Berkovitch) was not identical to the indenter tip used in this study (10pm hemi-spherical) the elastic response of the silicon is still expected to be negligible for indents made in the compliance specimens. Compliance data was also collected from indents made on the diamond surface of specimens 8, 11, 12, 14, and 21 in away from the diamond diaphragm. Loads during the compliance test on the faceted diamond surfaces were limited to less than 60mN in order to reduce the possibility of indenter tip damage. 60mN is beyond the range of maximum loads sustained by the diamond diaphragms before failure. 81 8.0 Off-center diaphragm loading To study the effect of off-center point loading of the diamond diaphragms, specimen 9 was loaded at various distances (0 to 400nm) from the center of the diaphragm. In order to prevent diaphragm failure before all the loadings could be performed, the maximum load was maintained at 30mN, which produced a central deflection of approximately 13pm. Figure 29 provides a schematic of the off- center loading of diaphragm specimen 9. 82 diamond diaphragm 400 um V r L v V 1570um Figure 29. Schematic of off-center loading of diaphragm specimen 9. 83 9.0 SEM, TEM, and X-ray analysis of diamond film After load-deflection data was collected for the diamond diaphragms, fragments of the diaphragms were collected for SEM' analysis. The collection of the diaphragm fragments was performed using two methods. One method involved the use of tweezers to fracture the diaphragm and collect the resulting diamond film fragments (Figure 30). The fragments were secured “on end” to a circular aluminum SEM stub coated with a low vapor pressure adhesive tab. The collection and handling of the diaphragm fragments with tweezers was very tedious and the diamond film fragments were often destroyed (broken into pieces too small to handle) during the process. A second, somewhat less tedious method, utilized a hammer to obtain fragments of the diaphragm for SEM analysis (Figure 31). The diaphragm specimen was placed on a table with the diaphragm side up. An aluminum SEM stub was placed on the specimen and centered over the diaphragm. A sharp hammer blow(s) to the aluminum stub resulted in the fracture of the silicon substrate and the debonding of the silicon substrate from the glass tube. The specimen tended to fracture into “pie shaped” segments and large pieces of the silicon substrate with portions of the diamond diaphragm still attached were created. The silicon/diaphragm fragments could be easily handled with tweezers and positioned on an aluminum SEM stub. Conductive (colloidal graphite) cement was used to secure the silicon/diamond fragments to the SEM stub. 84 Top View Of Al SEM Stub '\ adhesive tweezers fractured diaphragm diamond film fragments Figure 30. Diagram of “tweezer method” for the mounting of diamond film fragments for SEM examination. gianilond lap ra fragrnen'clzIsn \ “ Vt / graphite cement 1 Si substrate /7 tweezers Si substrate fragments Figure 31. Diagram of the “hammer method” for the mounting of diamond film diaphragm fragments for SEM examination. 85 Although this method is much less tedious than the first and often produced better results, it also resulted in an occasional catastrophic failure of the entire specimen. If the specimen did not fracture as expected, tweezers were then used to collect the diaphragm fragments. The aluminum SEM stubs with the diamond fragments were placed in an Emscope sputter coater to be coated with gold. The stubs were tilted slightly (~ 20°) in the sputter coater so as to allow the face of the fragments to be adequately coated with a layer of gold. After approximately 15nm (for horizontal surfaces) of gold had been deposited the specimens were tilted slightly in the opposite direction and another 15nm of gold was applied. The “double coating” of the specimens was performed to ensure adequate conductivity of the diamond fragments during SEM analysis. After gold coating, the stubs were placed in a JOEL35 SEM. Using the tilt and stage position controls, the diamond diaphragm fragments were located and oriented with the fracture surface perpendicular to the electron beam. Alignment of the diamond film fragment with the electron beam produced significant image distortion due to edge and other SEM specimen/beam interaction effects. The image distortion was used to locate the stage tilt angle at which a particular fragment was aligned with the beam; After alignment, the stage was tilted an additional 3° to improve image quality for film thickness measurements. The thickness of each fragment was measured at several locations directly from the CRT on the SEM using a plastic ruler and the 86 micron markers provided by the SEM. Several micrographs of the fracture surface of the diaphragm fragments were also taken. The average value of the fragment thickness measurements was used as the diamond diaphragm thickness. Prior to SEM analysis, a magnification calibration of the SEM was performed using a copper calibration disc (Ted Pella Associates), which is essentially a tiny ruler with 10pm line spacing. Any discrepancy in the indicated magnification was noted and used during the diaphragm thickness measurements. The working distance was found to be important with respect to the SEM magnification calibration. Specimen fragments to be analyzed were positioned at or very near the recommended working distance to ' insure accurate magnification indication. Specimens for TEM analysis were prepared from films obtained from wafer #1. Free-standing diamond film fragments (3mm x 3mm) were prepared by etching the substrate in 10 molar KOH at 60°C. ' The diamond fragments had a thickness of roughly 1.5mm and required thinning prior to TEM analysis. Thinning of the fragments was preformed using an argon ion mill. Analysis of the diamond film in a Phillips 80 TEM consisted of brightfield and axial darkfield imaging, as well as selected area diffraction (SAD). X-ray diffraction was performed on a diamond film obtained from wafer #1. The diamond film, still on the silicon substrate, was glued to the aluminum specimen holder of a Scintag XDSZOOO and a 20 scan of the specimen was performed. 87 10.0 Point load deflection of circular glass discs To test the ability of the nanoindenter and the load-deflection procedure for the measurement of elastic modulus, a glass disc specimen was prepared and tested. The circular glass disc specimens (Figure 32) were prepared using glass coverslips (thickness = .145- .210mm), superglue, and two rectangular glass plates (76mm x 51mm x 1mm). Circular holes (18mm and 23mm in diameter) were cut in the center of two glass plates using a diamond hole saw. An additional glass plate was superglued to the annular glass plates for added rigidity. The deflection of the glass plate was performed using the nanoindenter and a method similar to that used for diaphragm deflection. The glass plates could not be loaded to fracture without fear of damage to the nanoindenter. Maximum load for the glass diaphragms was 2.6 N. Two diaphragms were prepared by supergluing coverslips over the 18mm hole. Another diaphragm was prepared by placing (no glue) a coverslip over the 23mm hole in the glass plate. After deflecting the glass diaphragms with the nanoindenter, a compliance test was performed on the glass plates. The glass ' diaphragm sample was placed over the stage of the nanoindenter with the “backside” of the glass plate facing upward. Three indentations were made in the approximate center of the glass plate. Load-deflection data was collected by the nanoindenter and stored on a floppy disc. 88 76mm 51mm 7 m glass microscope slides glass cover slip surfirglue 1/ %e I 1.0mm —[ l I g I: 1 Figure 32. Diagram of glass diaphragm fabrication process. Specimens were prepared with and without glue around the rim. Bar—‘1‘:-_ " 89 III. RESULTS & DISCUSSION 11.0 Visual, SEM, TEM, and X-ray diffraction observations Actual diamond films obtained from wafer #1, and wafer #2, are presented in Figure 33 along with an additional diamond film which was used in LSCM buckling measurements but not in this diaphragm deflection study. The relative “whiteness” or clarity of the diamond film gives a rough indication of film quality (i.e. darker films are expected to contain a higher degree of graphitic or non-diamond carbons). Visual observations of film appearance indicate that the films used in this study, which are reasonably “white” were of fairly high quality. ' An optical stereographic microscope with magnifications up to 65x was also used to view the surfaces of the diamond films. Diaphragms obtained from wafer #1 (A-F) appeared to be free of foreign materials (i.e. particulates and residue picked up during processing) but a random distribution of relatively large (up to 30mm) particles were apparent on the faceted surface of diaphragms 6 -21 obtained from wafer #2. The particulates observed on diaphragms 6-21 resisted all attempts of removal, including rubbing, and when scraped with the edge of a razorblade (while being viewed under the microscope) considerable effort was required to dislodge them from the surface of the film. Figure 21 (Section 5) shows optical micrographs of typical diamond diaphragm specimens. All diamond diaphragms exhibited a 90 Figure 33. (A) Fragment of diamond film obtained from wafer #1 and corresponding SEM surface micrograph. (B) Fragment from wafer #2 and SEM surface micrograph. (C) Fragment of diamond film obtained from a wafer used in LSCM buckling analysis [122] but not in this deflection study. Included to illustrate differences in diamond film transparency and surface morphology. 91 high degree of circularity as observed visually, and based on measurements made using the nanoindenter, the diameters in any given direction varied by less than 5% for all diaphragms used in this experiment. ' SEM analysis of the top surface (surface away from the substrate) showed that the films consisted of highly faceted diamond crystallites (Figures 34-35). Specimens from wafer #1 had a surface dominated by (111) planes with twinned morphologies, while the surface of specimens obtained from wafer #2 had a large percentage of (100) as well as (111) planes. Surfaces of films obtained from both wafers is consistent with those of films reported by a number of researchers in the literature. A large number of smaller crystallites which do not appear to extend through the thickness of the film were present in specimens obtained from both wafers used in this study. These smaller crystallites are assumed to be the result of diamond nucleation on the growing diamond surfaces. The large particulates which were observed under an optical microscope on the surface of specimens obtained from wafer #2 were also observed in the SEM. As can be seen in Figures 36 and 37, the particles are covered with diamond crystallites, which means they were present before diamond deposition or were introduced during the deposition process. The nature of the large particulates is unknown and it is possible that they are made up entirely of diamond crystallites, but judging by their shape and size (up to 30pm) this seems unlikely. The presence of the particulates is not expected to 92 511m Figure 34. SEM surface micrograph of diamond specimen taken from wafer #1. 5pm Figure 35. SEM surface micrograph of diamond specimen taken from wafer #2. 93 significantly influence the deflection behavior of the diamond diaphragms. SEM observation of the back (substrate) side of the diamond films showed a relatively smooth surface with a number of large (up to ~0.2p.m x 211m) voids present between adjoining grains (Figure 38). The debris present on the substrate side of the films is assumed to be introduced during post deposition processing but the origin of the particles is not known. The grain size of the crystallites at the substrate surface of the films was approximately 1.2um for specimens obtained from wafer #1 and 111m for wafer #2. SEM imaging of the fracture surface of the diamond film diaphragms indicates the columnar structure typical of the CVD diamond growth process (Figure 39 - 41). The fracture of the diamond film appears to be inter-granular (along grain boundaries), as evident by the rough and uneven fracture surface. Figure 41 is a micrograph of a fractured diaphragm specimen taken in the region of the silicon rim. The silicon rim of the diaphragms gradually “tapers” down to the diamond film. This tapering is a result of the specimen preparation procedure (i.e. the grinding of a hemispherical dimple) and the effects of the tapered rim on the deflection behavior of the diaphragm will be discussed in upcoming sections. A diamond film specimen obtained from wafer #1 was analyzed using transmission electron microscopy. TEM analysis was intended to provide insight into the orientation relationship between the individual crystallites in the diamond films by using selected area diffraction to discern their relative orientations. Unfortunately, the 94 100nm Figure 36. SEM micrograph of specimen obtained from wafer#2 showing particles adhering to the surface of the diamond film. 10pm Figure 37. SEM micrograph of particulate. The diamond coating Indicates the particles are introduced before (or during) the deposition 95 lum Figure 38. SEM micrograph of back (substrate) surface of a diamond film specimen obtained from wafer #2. r"t , \ / . V. "'1" . .I ' ,3 'r r 5pm Figure 39. SEM micrograph of cleaved diamond film/ Si substrate obtained from specimen 13. 96 211m Figure 40. SEM micrograph of fracture surface of diamond diaphragm 16. 311m Figure 41. SEM micrograph of silicon rim supporting diamond diaphragm A showing the tapering of the silicon substrate to the edge of the diamond diaphragm. 97 double tilt holder for the TEM used in the study was not operational during the time the TEM analysis was performed, making the orientation of the individual diamond crystals virtually impossible to determine. Diffraction patterns were obtained for numerous diamond crystallites obtained using the smallest (30pm) SAD aperture on the TEM. Diffraction patterns indicative of a single crystal low-index zone (Figure 42) could only be obtained for three of the largest diamond crystallites (diameters ~ 3pm). Analysis of the diffraction patterns agreed with the theoretical predictions for a diamond <110> zone for each of the three diamond crystals from which they were obtained. Streaking and extra spots around the diffraction spots corresponding to the <111> direction indicates twinning parallel to {111} planes. Spot splitting in the diffraction patterns was also observed and is attributed to the presence of stacking faults in the diamond crystallites. Axial dark-field images using diffraction Spots from the diamond <110> zone patterns provided no evidence of any orientation relationship between the surrounding crystallites and the crystal from which the low-index diffraction pattern was obtained. It is interesting to note that the only low index zone pattern produced was that of the <110> zone, which agrees with the <110> texturing commonly observed in CVD diamond films. The TEM analysis of this study, however, is far too limited to make any conclusions as to the nature of the texturing (if 3113') in the diamond film studied. Figure 43 is the result of an X-ray 20 scan'of the diamond film obtained from wafer #1. The relatively small diamond peaks, and the 98 Figure 42. TEM selected area diffraction pattern taken from a single diamond grain from a thin film specimen obtained from wafer #1. The pattern agrees with that of the theoretical <110> zone for a diamond crystal. Streaking and spot splitting indicate the presence of twins and stacking faults in the diamond grain. Spots resulting from double diffraction are also present. 300 lllllllllllllllillillllIlllllllLLlLMll'lllillllill X-RAY 26 SCAN OF DIAMOND .FILM ON SILICON SUBSTRATE s'°°"“°°’ 200— — .‘2 C 3 O O 100~ - Diamond(111) Diammdtzzof c IITIIIIIIIIiTjTIIIIIIIl[ITIITITIIIIIIIIIIYYITIIII 30 35 4O 45 50 55 60 65 70 75 80 29 Figure 43. X-ray 20 scan of diamond film on silicon substrate obtained from wafer #1. 99 large (100) silicon peak is due to the low mass absorption and relative “thinness” (~1.5um) of the diamond film. The diamond (111) and (400) peaks are readily distinguishable and occur at the theoretical Bragg angles for diamond. 100 12.0 Diaphragm morphology - buckling relief of stress. Visual observations of the diamond diaphragms A - F indicated a “blistered and wrinkled” morphology. Visual observation of diaphragms 6 -21 however, gave no indication of any deviation from “flat”. The apparent shape of the buckled diamond diaphragms (A - F) is consistent with observations of Argon et al [96] on CVD silicon carbide films in biaxial compression, grown on single crystal silicon substrates. The SiC films used in the study by Argon buckled upon release from the substrate. The morphology of the delaminated circular SiC “blisters” consisted of a single mode diametral buckling combined with multi-modal buckling around the circumference. The buckling of the SiC blisters is attributed to the release of biaxial residual compressive stress in the film. Upon buckling, Argon et al. assumed the residual compressive stresses in the SiC were almost completely relieved [96]. The buckling behavior observed in the diamond diaphragms used in this study is also assumed to result from the relief of biaxial compressive residual stress. Based on the buckling criterion for a clamped circular plate with a uniform in-plane load around the circumference [124] and the relationship between force/unit edge length and residual stress, the biaxial compressive stress required to buckle the diamond diaphragms used in this study was calculated (Equation 12.1). NCr = critical force/unit edge length for buckling D t D 2 plate rigidity (Equation 3.3) oer = Ncr t = 14.68 72" r = plate radius (12.1) t = plate thickness on = critical residual compressive stress , 101 For diaphragms A - F a residual compressive stress of 5MPa as calculated from Equation 12.1 is sufficient to buckle the diamond diaphragms. For specimens 6-12, Equation 12.1 predicts a residual compressive stress of 15MPa is not enough to produce buckling in any of the specimens and a compressive stress of approximately 40MPa is required to buckle the most rigid diaphragm used in this study. Although 40MPa is a fairly substantial stress, it is an order of magnitude less than the calculated compressive stress of approximately 500 MPa resulting from thermal expansion mismatch between the diamond film and substrate (Section 2.1). Due to intrinsic factors the net residual stress in the diamond film may be less (or more) than the thermally induced stress, but the presence of a residual compressive stress in the diamond films sufficient to buckle the diaphragms is very likely. Although the rigidity of the diamond films used in this study is significantly higher than that of the SiC films in the Argon study and the calculated compressive stress is lower, it is assumed that the majority of the residual compressive stress in the diamond diaphragm is relieved upon buckling. The compressive stress in the region around the rim of the diaphragm can not be relieved by buckling due to the constraint imposed by the silicon substrate. The circumferential buckling in the diaphragm was not observed in the region close to the rim of the diaphragm due to the “clamping” of the diamond film by the substrate. Circumferential buckling in the diaphragm was evident in regions less than 50pm from the rim of the diamond diaphragm. The maximum magnitude of the 102 circumferential buckles as measured by the LSCM was approximately 311m and occurred at approximately one-third of the distance between the rim and center of the diaphragm. Circumferential buckles in diaphragms obtained from wafer #1 were visually observed using the naked eye or a stereographic optical microscope by holding the specimen at a low angle with respect to the light source. The number of circumferential buckles observed around the rim of the diaphragms (from wafer #1) was typically 8 to 12 and no correlation between diaphragm size and the number of circumferential buckles was apparent for diaphragms with diameters ranging from 1200 - 2000um and a thickness of approximately 1.5 um. Circumferential buckling was not visually observed in diaphragms obtained from wafer #2. Argon et al. [96] used the wavelength of the buckled SiC blisters to predict the residual stress in the film. Due to the nature of the “delamination front” of the growing blister and the relatively low rigidity of the SiC thin film, the number circumferential buckles in the SiC blister increased with increasing diameter. Blisters up to approximately 150nm in diameter and film thickness ranging from approximately 0.1 to lum were used in the Argon et al. study. The diameter of the diamond diaphragms, ranging from 1200 - 2000 pm in this study, was defined by the grinding and etching process. Based on the lack of correlation between the diaphragm diameter and the number of observed circumferentialbuckles, it is believed that the number of circumferential buckles which first initiate in the diaphragm during the etching process is maintained 103 as the diameter of the diaphragm increases. The “steady state” of circumferential buckles observed in specimens obtained from wafer #1 may result from the magnitude of residual stress in the diamond films being insufficient to overcome the diaphragm rigidity and produce a new buckling wavefront for the range of diaphragm diameters observed. The method used by Argon et al. to determine the residual stress in thin films using the wavelength of circumferential buckling does not seem to be applicable to the diamond diaphragms used in this study due to the apparent “steady state” of the diamond diaphragm circumferential buckling with increasing diaphragm diameter. Observations of the buckling morphology of diamond diaphragms with larger diameters used in an associated study by Flowers [125] showed a complex “rippled” morphology which may result from the linkup of opposing circumferential buckles producing a multi-modal morphology across the diameter. The complex morphology of the diamond films with larger diameters (>3um) would make the modeling of the deflection behavior much more difficult, thus, diaphragm diameters were maintained between 1 and 2mm. The “rippling” of the films was not observed in the study by Flowers for diaphragms with diameters of 2pm or less. The magnitude of the single mode diametral buckling in several diaphragms measured by the LSCM technique (all from wafer #1) was related to the diaphragm diameter (i.e. larger diaphragms had greater central deflections due to buckling). LSCM measurement of central deflections on several diaphragms of various diameters (1200 104 - 2000) indicated that the magnitude of the initial (stress induced) central deflection was roughly 1% of the diaphragm diameter. The LSCM measurements were also consistent with observations of initial central deflections made with the nanoindenter (Section 7 .2). The majority of the diaphragms which exhibited observable buckling were “convex” with respect to the faceted surface of the diamond film after buckling, but a few were observed to be “concave”. The direction (concave or convex) of buckling had no observable effect on the morphology of the diaphragm. In the LSCM line scan analysis of diamond diaphragm specimen A (Section 6), the average value for the measured strain (Equation 6.1) in specimen A induced by the buckling of the diaphragm to relieve the compressive residual stress in the film was ~ 5.6 x 10"4 (Table 6). The average measured strain induced in the diaphragm due to buckling in specimen A is very Similar to the amount of strain due to thermal expansion mismatch predicted from thermal expansion polynomials (Figure 10). However, the residual stress in the film is not completely relieved by buckling of the diaphragm and the net stress in the film is influenced by intrinsic factors as well as thermal expansion mismatch. The calculated compressive stress in specimen A is fairly large (>500MPa) as determined by multiplying the measured buckling strain relief by the modulus of natural diamond. The similarity between the measured strain relief and the estimated thermally induced strain implies that the net intrinsic stress in diaphragm A is relatively small, and depending upon the degree of strain relieved by the buckling of the diaphragm, the intrinsic stress 105 may be compressive. Differences in CVD diamond film thickness at different points across a single substrate are commonly observed in CVD diamond films which indicates that growth conditions can vary with wafer location. It is therefore conceivable that significant differences in the residual stress of the diamond film diaphragms with respect to the relative wafer position may exist. Specimens A-F were obtained from wafer #1 and had similar thickness (within 10%) and surface morphologies, thus they are expected to have similar residual stress states. Specimens 6-21 were obtained from wafer #2 and had similar surface morphologies but diamond film thickness between the individual diaphragms varied by as much as 40%. Therefore, the intrinsic stress in diaphragms 6-21 may vary significantly. 106 13.0 Deflection behavior of point loaded circular glass discs The circular glass diaphragms, described in Section 9, were used as a rough calibration of the nanoindenter and a test of the diaphragm deflection method in measuring the elastic modulus of a material. The results for the deflection of the glass diaphragms are presented in Figures 44-45 The relatively large thickness (145 - 210nm) and large deflection (up to 27 011m) of the glass diaphragms allowed analysis of the deflection behavior both by classical (linear) and nonlinear plate theories (Section 4). The load-deflection curves for the simply supported and clamped glass discs were fit to the approximate nonlinear solutions for a point loaded circular plate (Equation 4.2) derived by Timoshenko et al. and Banerjee et al. (Section 4). The fitting produced a value for the elastic modulus (E) and the constant A describing the nonlinear behavior of the plate (Section 4, Table 4). Fitting the simply supported circular glass plate with the general form of the approximate plate deflection solution (Section 4, Table 4); E 2 fit elastic modulus Et4 ] { [W] [WP } A = fit constant (Table 4) [—-—1'2 "' + A — = P w/t = normalized deflection (13. I) '551 t t t = plate thickness r = plate radius resulted in: E = 46 GPa, A = .20, correlation = .999. The value of .20 obtained for A agrees with the values of A derived by Bannerjee and Timoshenko (.16 and .27 respectively), but the fit modulus of 46 GPa is substantially lower than the expected modulus of the coverslip deflection (microns) 107 300 m 1 1 __L L l 1 I . l g 1 J 1 4 1 SIMPLY-SUPPORTED COVERSLIP support radius = 11.1mm I" ' coverslb radius = 12.7mm m:- 250 -' thickness 3 .1450?" ”1131"" _ .1ifttl9'ji!"w 200 _ .1c.l""""" ' _ 150 _ ".....'..:.‘§:":’ J 100 - .. 50 - _ ' compliance 0 T I w 1 fi 1 r r l v I v ]— V . I l— 0 200 400 600 800 1000 1200 1400 load mm 1 ' I ‘ 1600 1800 2000 Figure 44. Load-deflection behavior of simply-supported circular glass coverslip prepared by placing (no glue) the coverslip over a circular hole in a glass plate. The compliance curve obtained from indentation into the back of the glass plate(s) was subtracted from the data obtained for the clamped and simply supported plates. 108 150 CLAMPED CI'RCULAR' GLASS ' OVERSLIPS deflection (microns) L3 radius = 9.0 mm 1 = 145 microns t=210 microns W 1500 load (mN) 500 1000 2000 I 2500 3000 Figure 45. Load-deflection plots obtained from point loading of circular glass coverslips clamped by applying superglue around the rim. 109 (Eglass ~ 70 GPa). Fitting glass plate #2 (clamped) using; A = fit constant of nonlinearity (Table 4) [gfiilé‘z] { [‘11] + A it! 3} = P E/::ii:l::fi:§ddlelfil:ction ( 13.2) t 2 plate thickness r = plate radius provided an elastic modulus (E) of 53 GPa, A = .19, and a correlation coefficient of 0.999. The solutions obtained by Timoshenko and Banerjee (Section 4) predict a value of ~.43 for the constant (A) describing the nonlinearity of the load-deflection curve for a fully clamped plate, which is over twice that obtained by the fitting. The fit modulus of 53 GPa is also substantially lower than the expected modulus of the glass coverslip (7 0GPa). Loading clamped coverslip #3 (t = 210nm, r=9000um) to 1.9N (the maximum load allowed by the nanoindenter for automatic data collection) produced a central deflection of approximately 1/3 the plate thickness, which is not enough to observe significant nonlinear behavior. Classical (linear) plate theory (Section 4, Equation 4.1) for a point loaded clamped circular plate was used to estimate the modulus of glass coverslip #3 (Equation 13.3) r = plate radius E = elastic modulus 1V = 312(1-V2) v = Poissons ratio (13 3) P 41tEt3 t = plate thickness ° wlp = slope of load-deflection curve (4.0 x 10'5 (m/N) resulted in an estimated modulus (E) of 50 GPa. The modulus obtained for clamped coverslip #3 using classical (Equation 4.1) plate theory (50GPa) is similar to the modulus 110 predicted using nonlinear theory (Equation 4.2) on both the simply supported coverslip #1 (46 GPa) and clamped coverslip #2 (53GPa). For the coverslips which were superglued around the rim the criterion for a fully clamped plate may not have been met, resulting in larger deflections than predicted by theory (Equation 4.2) and a lower estimated modulus. The value of A, the constant describing the nonlinearity of the plate, fit to coverslip #2 (.19) was less than one-half the value given by the approximate solutions (Section 4, Table 4) for a fully clamped plate. The large difference in the theoretical and fit value of A may indicate that the superglued coverslips were not fully clamped. The coverslip which was not glued around the rim was presumed to be simply supported with the edge free to move. The value of A obtained from the fitting agreed with the theoretical value of A (Section 4, Table 4) for a point loaded simply supported circular plate with a moveable edge. For the simply supported case, boundary conditions would not appear to be the cause of the seemingly low estimate for the elastic modulus of the glass coverslip. Frictional forces around the rim in the simply supported case may give rise to an inflated measured modulus value. Possible explanations for the “lower than expected” measured modulus for the glass coverslips was that the balance on the nanoindenter was improperly calibrated or that the displacements reported by the nanoindenter were not the displacements being induced. Both of these possibilities were investigated using the nanoindenter calibration procedure described in Appendix B. The results of the nanoindenter calibration (Appendix B) indicate the 111 displacements induced by the nanoindenter are 66% of the reported displacement. Appropriate adjustment to the measured modulus values of the glass coverslips produces values of 69, 79 and 75 GPa for coverslips #1, #2, and #3 respectively. Only calibrated values for nanoindenter displacement will be reported in the remainder of this thesis. The load-deflection behavior of the clamped circular coverslips was normalized (Equation 13.4) following the relationship between point load and thickness predicted by plate theory (Equation 4.1,4.2). P P = point load 73' t = plate thickness (13.4) normalized load = The load-deflection curves after normalization of the load for the - two coverslips (t = 210 and 145 um) clamped by gluing are depicted in Figure 47. The normalized load-deflection behavior for clamped coverslips #2 and #3 roughly coincide. The difference in the curves after normalization may be the result of differences in the effective clamping of the glass coverslips. 112 lCLAMPED CIRCULAR GLASS COVERSLIPS WITH LOAD NORMALIZED . '8 m- 2 0 ° .9 o S 0 ° .5 D ‘6 g ‘3 radius = 9.0 m D t = 145 microns ‘ 0 t=210 microns 1 . , . . , . , . .0002 .0004 .0006 .0008 .001 normalized load (mN/ microns ) Figure 47. Load-deflection plots for clamped glass coverslips with load normalized for radius and thickness. 113 14.0 Deflection behavior of point loaded diamond diaphragms Nanoindenter-specimen compliance data collected from the various compliance specimens (Section 7.4) was combined into a single data set (Figure 48). The compliance behavior was described using Equation14.1. The form of Equation 14.1 was chosen by the author to describe the saturation to linear behavior observed inthe compliance data. The constants in equation 14.1 were obtained by manually fitting to the entire set of compliance data. d = 0.01P + 0.4 [1 - exp(-P('6)] 3:333:31“ (14.1) Equation 14.1 was used to subtract the nanoindenter-specimen compliance contribution from the load-deflection data for the diamond diaphragms. The relative contribution of the nanoindenter- specimen compliance to the total deflection of the diamond diaphragms is fairly small (<5%) for the large load (>25mN) deflection behavior of the diamond diaphragms. At low loads (<5mM) the contribution of the compliance to the total deflection measurement is fairly substantial (~25%). The diameter, thickness, load at failure, and number of data points collected for the deflection of the diamond diaphragms used in this study are presented in Table 7. The relative position of diamond film diaphragm specimens 6 - 21 with respect to the silicon wafer on which they were deposited is presented in Figure 15 (Section 5). The measured thickness of the diaphragm specimens corresponds to the 114 displacement (microns) load (mN) Figure 48. Plot of combined compliance curves taken from various diamond specimens and specially prepared compliance specimens. The compliance behavior was described by Equation 14.1 which was used to subtract the compliance contribution from the diaphragm deflection data. 115 Table 7: Radius, thickness, and load at fracture for point loaded circular diamond diaphragm specimens. specimen radius (mm) thickness (um) fracture load (mN) A 648 1.5 5.8 B 719 1.55 3.4 C 868 1.5 4.9 D 976 1.6 2.8 E 866 1.5 1.9 F 907 1.65 4.6 6 775 4.0 42 8 875 4.4 ' 40 9 785 4.3 31 11 890 3.1 17 12 962 4.3 36 13 865 4.3 37 14 835 4.1 37 16 895 4.0 30 17 670 3.8 32 19 905 3,5 27 21 900 3.2 ' 21 pl dl. dt 116 expected trend of thickness variation with respect to the position on the wafer from which the specimen was obtained (i.e. thickness decreases as distance from the center of the wafer increases). Film thickness near the center of the wafer is expected to be greater due to its position relative to the plasma ball generated during the MPACVD process (Section 1.4). The portions of the wafer which were not used in the diaphragm deflection experiments consisted of specimens which were either destroyed during processing or fragments which were unsuitable for the production of a diamond diaphragm. The load-deflection data for the circular diamond diaphragms, corrected for nanoindenter-specimen compliance, was fit using a non- linear least squares regression program, written by E. Case, to Equation 14.2. Equation 14.2 is the form used by the approximate solutions of Timoshenko et al. and Bannerjee et al. to describe the large deflection behavior of circular plates subjected to a central point load (Section 4). Prior to fitting, the deflection data was normalized (divided by the diaphragm thickness) in order to comply with the general form of Equation 14.2. P = point load C1 (X + C2X3) = P Cl, CZ = fitting parameters (142) X = w/t (normalized deflection) Results of the load-deflection data fitting is presented in Table 8 and plots of the load deflection data and respective fits for diamond diaphragms A - 21 are shown in Figures 49-66. Equation 14.2 described the load-deflection behavior of every one of the 17 diamond w/t Figu 117 SPECIMEN A thickness =1.5 urn ‘ radius = 640 P (mN) Figure 49. Load-deflection plot for diamond diaphragm specimen A. 118 SPECIMEN B thickness = 1.55 pm ‘ radius = 719 —4 ‘ P (mN) Figure 50. Load-deflection plot for diamond diaphragm specimen B. 119 SPECIMEN c ‘ A 0 thickness =1.5 I4 m ‘ radius = 868 ” -1 P (mN) Figure 51. Load-deflection plot for diamond diaphragm specimen C. 120 SPECIMEN o . thickness =1.6 “m radius = 976 c—i — A P (mN) Figure 52. Load-deflection plot for diamond diaphragm specimen D. 121 SPECIMEN E thickness =1.5 um * radius = 866 —: --4 — Figure 53. Load-deflection plot for diamond diaphragm specimen E. 122 SPECIMEN F thickness = 1.65 gm ‘ radius = 907 —l P (mN) Figure 54. Load-deflection plot for diamond diaphragm specimen F. 123 4 . l l 1 l SPECIMEN 6 P thickness = 4.0 “m radius = 775 3 - .. w/t 0 V I V ’l— 1 ‘I’ 1 I 0 10 20 30 40 load (mN) Figure 55. Load-deflection plot for diamond diaphragm specimen 6. Deflection to 9pm repeated three times. 124 4 4 1 l J l SPECIMEN 8 _ thickness = 4.4 um radus = 875 3 .. w/t o . 4 . T . f - I 0 10 20 30 40 load (mN) Figure 56. Load-deflection plot for diamond diaphragm specimen 8. Deflection to 9um repeated three times. 125 4 1 l l l I SPECIMEN 9 _ thickness = 4.3 ,um radus = 785 3 _ < 2 _ 3 1 _ 0 v I ' T . I ' 1 O 10 20 30 40 load mm Figure 57. Load-deflection plot for diamond diaphragm specimen 9. Deflection to 9pm repeated two times. 126 SPECIMEN L11 Fthickness = 3.1 ,um radius = 890 l i s- / - f r r T y I 20 30 40 load (mN) Figure 58. Load-deflection plot for diamond diaphragm specimen 11. Deflection to 9pm repeated three times. 127 4 x A l L 1 l SPECIMEN 12 _ thickness = 4.3 Hm radus =962 3 .. r / K" 3 0 ' I v I ' 7 . . I . 0 10 20 30 40 load (mN) Figure 59. Load-deflection plot for diamond diaphragm specimen 12. Deflection to 9pm repeated three times. 128 4 . l l I l SPECIMEN 13 _ “Ckness 3 4.3 “m radus = 865 3.. w/t O ' I . I ' F v I O 10 20 3O 40 load (mN) Figure 60. Load-deflection plot for diamond diaphragm specimen 13. Deflection to 9pm repeated three times. 129 A l 4 4 A l SPECIMEN 14 thickness = 4.1 um F racius = 835 w/t 0 . r - , . ' T 1 , 0 10 20 30 40 load (mN) Figure 61. Load-deflection plot for diamond diaphragm specimen 14. Deflection to 9pm repeated three times. 130 SPECIMEN 16 _ thickness = 4.0 pm diameter 8 895 w/t o r I V I Y f Y I 0 10 20 30 40 load (mN) Figure 62. Load-deflection plot for diamond diaphragm specimen 16. Deflection to 9pm repeated four times. 131 4 - 4 1 A 1 1 SPECIMEN 'I7 _ mass = 3.8 “m radus = 670 3 r 7 g 4 0 . I . I 0 _ 10 20 load (mN) 30 40 Figure 63. Load-deflection plot for diamond diaphragm specimen 17. Deflection to 9pm repeated three times. 132 SPECIMEN 19 _ thickness = 3.5 “m racius =905 O f T w I * I * V 0 10 20 30 40 load (mN) Figure 64. Load-deflection plot for diamond diaphragm specimen 19. Deflection to 9pm repeated three times. 133 SPECIMEN 21 thickness = 32 um radus =900 I o ' i * T ' V j I 0 10 20 30 40 load (mN) Figure 65. Load-deflection plot for diamond diaphragm specimen 21. Deflection to 9pm repeated two times. 134 diaphragms included in this study with a correlation coefficient in excess of 0.995. The nonlinear behavior exhibited by the diamond diaphragms during loading follows the general trend predicted by plate theory (Equation 4.2). During the early stages of the deflection, many of the diaphragms behaved in a manner which contradicts the predictions of plate theory (i.e. further deflection of the plate required a less than proportional increase'in load). The load-deflection behavior of specimen A (Figure 49) shows a change in curvature between the small (less than the thickness) deflection and large deflection behavior. The small deflection behavior is believed to result from the transition of the diaphragm morphology from that of a plate buckled by a uniform load around its circumference (Section 11), to the morphology of a plate deflected by a central point load. The point loading of a buckled diaphragm may also have a “smoothing” effect on the wrinkled morphology observed in diaphragms A - F and expected for diaphragms 6 - 21. In any case, the initial point loading deflection behavior is not expected to be completely explained by the general form of nonlinear plate theory (Equation 4.2) which describes the behavior of plates which are initially smooth and flat. The constant C1 obtained from the fitting of the load-deflection data for the diamond diaphragm specimens (Table 8) is predicted by non-linear plate theory (Equation 4.2) to have the following relationship to the diaphragm radius and thickness: C1 = fit constant to diaphragm load-deflection data. [ t4 ] r = diaphragm radius C 1 = (l :2 t = diaphragm thickness (143) a = constant function of Elastic modulus and Poissons ratio 135 Table 8: Specimen parameters and constant values for fitting of load-deflection data to Equation 14.2*. specimen 1.1133111: thiifigffs 1333:; C 1 CZ R A 648 1.5 50 1.045 .020 .998 B 719 1.55 27 1.038 -.006 .999 C 868 1.5 40 .492 .009 .998 D 976 1.6 19 .725 -.016 .997 E 866 1.5 10 .588 -.003 .999 F 907 1.65 37 .693 .017 .999 6 775 4.0 946 7,924 .022 .999 8 875 4.4 622 8.008 .017 .998 9 785 4.3 598 9,701 . .021 .999 11 890 3.1 922 1.821 .040 .998 12 962 4.3 928 7.443 .015 .999 13 865 4.3 927 8.338 .016 .997 14 835 4.1 1268 5.850 .019 .999 16 895 4.0 1095 5.840 .019 .998 17 670 3.8 1147 7.232 .019 .999 19 905 3.5 1094 4.015 .015 .999 21 900 3.2 754 2.698 .024 .999 * Equation 14.2: C1[ Viv + C2{%v }3] = P w = central deflection t = diaphragm thickness P = point load l 136 l ,(t‘ /r2 I vs. CI specimens 8 - 21 (microns2 x 104) 2 5 thickness? Iradiu Figure 66. Plot of C1 (Equation 14.3) versus t4/r2 for all specimens. Dashed line represents fit forced through zero, Solid line represents best fit with floating intercept. 137 The results of plotting C1 versus t‘I/r2 for specimens sets 6-21 and A- F respectively are presented in Figure 66. The predicted linear relationship (Equation 14.3) between fit constant C1 and t4/r2 for the point-loaded diamond diaphragms is observed with a correlation of .92 (force through zero) and .96 (not forced through zero). The initial (residual stress induced) central deflection of the diaphragms as well as differences in the effective diaphragm boundary conditions may account for the scatter in the C1 versus t"‘/r2 data. Specimens A-F were loaded on the substrate side and thus the faceted side of the film is were the maximum tensile stress will occur. The rough faceted surface may act to produce areas of highly concentrated stress which would result in a significantly lower fracture stress than a diaphragm loaded on the faceted side (6-21). The relatively large difference between the behavior of the grouped specimens (6-21 and A-F.) may also be explained by significant differences in initial , (residual stress induced) central deflection and boundry conditions. Specimens A-F and 6-21 were obtained from silicon-diamond wafer#1 and wafer#2 respectively, which were grown under differing CVD conditions and have different surface morphologies. Therefore, the possibility of significant differences in residual stress between the two specimen sets exists and would result in differing respective compressive stress induced diaphragm deflections. Specimens A-F had the point load applied to the substrate side of the diamond film and were supported from above by the silicon substrate (i.e. diamond film adhesion to the silicon substrate was the sole means of diaphragm support). Specimens 6-21 were loaded on the faceted side 138 of the diamond film and were supported from below by the silicon substrate. The silicon rim around diaphragm specimens A-F was not as drastically “tapered” as the rim supporting specimens 6-21, which may also effect the relative diaphragm boundary conditions. Specimens 6-21 were roughly 2 to 3 times the thickness of specimens A-F and therefore have a rigidity (Equation 3.3) approximately 1 to 2 orders of magnitude higher. The larger relative rigidity may make the buckling behavior of specimens 6-21 different than that of specimens A-F. 14.1 Repeated loading of diamond diaphragms To evaluate the repeatability of the diaphragm deflection behavior and address the possibility of smoothing, delamination, slippage, or other effects which would alter the response of the point loaded diaphragm, several specimens (6 - 19) were subjected to repeated loadings. Loading to 911m central deflection, unloading, then reloading to 911m, a procedure which was repeated up to 4 times, did not alter the load / deflection behavior by more than 5%. Due to the nature of the nanoindenter’s stage motion mechanism (gears), the vertical location at which the indenter contacted the diaphragm upon loading could not be accurately compared to the position at which, upon unloading, they became separated. The repeated loadings (Figures 6-21) of the individual diamond diaphragms were typically indistinguishable and in the worst case (specimen 13) had a maximum difference in load value for a given displacement of < 10%. The conformity of the repeated diaphragm loading behavior indicates 139 that the effect of smoothing, delaminations, etc. either do not occur, or have little effect on the deflection behavior. 14.2 Off-center diaphragm loading The effect of off-center loading on the diaphragm load-deflection behavior was investigated (Section 8). Diaphragm specimen 9 was loaded at several points away from the center and the resulting curves are depicted in Figure 67. The trend of the load-deflection curves (i.e. increasing the distance from the center decreases the deflection for a given load value) agrees with the expected load- deflection behavior (i.e. as the distance from the center increases, larger loads are necessary to induce a deflection equivalent to that of the centrally loaded case). It is also evident from the grouping of the load-deflection curves for loadings at distances less than 100nm from the center, that the load-deflection behavior of the diamond diaphragm will not be significantly affected if the center is missed by even a substantial amount (100nm or roughly 7% of the diaphragm diameter). The method used in this study to locate the center of the diamond diaphragms is accurate to 100nm. 140 ‘n I————l—__I_— _ l I l I _OFF CENTER LOADING 9 - _ 8 - _ 7 — _ ’6 L C 5 6 " ‘ E * DISPLACENENT‘ " 5- FROM CENTER- g (microns) % o centered ‘ 8 4 o as - 8 . 50 3 o 75 _ v 100 n 150 2 c 200 ‘ 3 250 1 ' 300 - l 350 A 400 0 I 4 T 0 10 20 30 load (mN) Figure 67. Load-deflection plots for off-center point loading of diamond diaphragm specimen 9. ' 141 15.0 Analysis of load at failure for the diamond diaphragms Timoshenko et al. [113] predicts the maximum tensile stress in a point loaded circular diaphragm will occur at the lower surface centered directly beneath the point load. Compressive stresses in the vicinity of the point load contact may exceed the maximum tensile stress in the diaphragm. The strength of brittle materials (i.e. diamond) in compression is typically several times larger than the tensile strength [113]. Therefore, the diamond films are expected to fail as a result of tensile stress. Timoshenko et al. describes the relationship between maximum tensile stress in a fully clamped plate (omax) and a highly concentrated point load (P) with the following equation (Equation 15.1): om,x = maximum tensile stress P P = concentrated point load Gmax = ?(1+V)[.485 111% + 0.52] tiplztie thlfklletses (15.1) r — ra us 0 p a n = Poissons ratio A highly concentrated load is defined by the ratio of contact area (c) and plate thickness (t) being small (~0). The point load applied to the diaphragms using a hemi-spherical diamond indenter with a 10nm radius is not expected to apply a “highly concentrated load” to the diamond diaphragms (t = 1.5 - 4.4 um) and may invalidate the use of Equation 15.1 in determining the maximum tensile stress at failure. The buckling of the diaphragms also violates the “flat plate” assumption of Equation 15.1. Due to the aforementioned factors, Equation 15.1 is not expected to provide an accurate estimate of the maximum tensile stress in the diamond diaphragms. The relation- 142 ship between thickness (t2) to the maximum tensile stress (omax) is observed in fracture behavior of the diamond thin film diaphragms (Figure 68). 143 50 1 I 1 m to L = 012 _ L = load at fallue t = fim thickness C = 2.05 40- R = .97 _ [I Load at iaflure (mN) 0 I I 0 10 film thickness2 (microns) 2 Figure 68. Plot of load at diaphragm failure versus t2. 144 16.0 Parameters influencing the diaphragm deflection behavior The load-deflection behavior of the diamond diaphragms used in this study is influenced by boundary conditions, initial (stress induced) deflections, as well as the buckled morphology of the diaphragm (i.e. the presence of wrinkles, etc.) [130]. The tapered silicon boundary (Figure 41) around the periphery of the diamond diaphragm is not expected to rigidly clamp the diaphragm during loading. Thus, the diaphragm boundary conditions do not conform to those. specified by established plate theory (i.e. fully clamped). The use of clamped plate theory to describe the behavior of a “sort-of-clamped” diaphragm may bring about significant error. However, the magnitude of the errors introduced by the non-idealized boundry conditions was not evaluated in this thesis and the author is not aware of an analytical study that directly addresses this problem. Residual stress in the diamond film is also expected to influence the load-deflection behavior of the diaphragms. This study provided insight into the buckling behavior of diaphragms formed from diamond films in residual compression. The degree of residual stress in the films, both before and after buckling, was not measured in this study. The effect of residual stresses remaining in the diaphragm after buckling (if any) on the diaphragm load-deflection behavior is unknown. 145 IV. CONCLUSIONS Circular diamond film diaphragms were prepared using a mechanical grinding and chemical etching process. Upon release from the substrate, buckling was observed in diamond diaphragms A- F obtained from a single silicon wafer (wafer #1) with film thickness less than 1.65 pm. Buckling was not visually observed in diaphragms 6-21 obtained from wafer #2 with film thickness, in excess of 3.1 pm (refer to Table 7). Diaphragm buckling was attributed to the presence of residual compressive stress in the film. Laser scanning confocal microscopy (LSCM) was used to analyze the morphology of the buckled diaphragms from wafer #1 and estimate the net residual strain in the films at room temperature. The LSCM measurement of net residual strain induced in the diamond diaphragm (5.6 x 10'“), in combination with the estimated strain at room temperature induced by thermal‘expansion mismatch between the silicon substrate and diamond film (~5.5 x 10'4), allowed for speculation into the nature of the intrinsic, or deposition induced, residual stress in the diamond films. Centrally point loading the diamond diaphragms with thickness ranging from 1.5 pm to 4.4 pm and diameters between 1275 um and 1925 um using a nanoindenter provided load-deflection data. Analysis of the diaphragms based on established approximate solutions to the point load deflection of circular plates was performed. Plate theory (Section 4, Equation 4.2) predicts a geometrically nonlinear behavior during point loading, which was 146 observed in the load-deflection behavior of the diamond diaphragms. The load-deflection behavior of the 17 diamond diaphragms used in this study was described by the form of Equation 4.2: r = radius of plate E = plate elastic modulus 3. P12 t = plate thickness (N + AN )= B[-—4] P = point load (4.2) Et N = normalized deflection (w/t) A,B = constants (Table 4) with a correlation coefficient in excess of .995. The effect of thickness and radius on the diaphragm deflection behavior which was predicted by established plate theory (Equation 14.3) was also observed in this study. The fracture behavior of the films also follows the predicted (Equation 15.1) diaphragm thickness dependance. Unconventional boundary conditions and initial (residual stress induced) deflections, however, did not allow the point-load behavior of the diamond diaphragms to be effectively modeled using flat plate solutions for idealized boundary conditions. Suggestions for the improvement of the process and the potential of the LSCM and point-load deflection techniques for the study of polycrystalline diamond films are discussed. 147 V. FUTURE STUDIES The magnitude and nature of the intrinsic stress for diamond films (such as the diaphragms used in this study) needs to be determined accurately. Future research utilizing the buckling morphology, measurements at various temperatures, as well as models for the degree of buckling strain relief may yield more detailed information on the nature of intrinsic stress in polycrystalline diamond films. The initial, residual stress induced, deflection of the diamond diaphragms is a significant problem to the point-load deflection analysis. Due to buckling upon release from the substrate, the diaphragms were not “flat plates” necessary to effectively utilize the established plate deflection solutions. Instead, the diaphragms assume a fairly complex “shell-like” morphology, to which solutions for point-load deflection behavior are not currently available and would be very difficult to develop. Again, a possible solution to this problem would be through sample preparation. Due to the complex residual stress state apparent in diamond films, the preparation of “flat”, stress free diaphragms may prove challenging, but seems very plausible. Boundary conditions for the diaphragms used in this study do not cOnform to those for which established plate deflection solutions exist. It may be possible to develop a plate deflection solution for the boundary conditions possessed by the diamond diaphragms in this study. If boundary conditions were the only obstacle facing the point 148 load analysis of the diamond diaphragms, development of a model for the actual diaphragm boundary conditions would be a worthwhile effort. Such a model would be difficult to test experimentally and would therefore be highly questionable. Another solution to the problem of unconventional diaphragm boundary conditions is through specimen preparation. Changing the specimen preparation procedure to produce diamond diaphragms with boundary conditions that conform to established plate theories would greatly improve the results of a similar study. The fracture behavior of the diaphragm was not fully analyzed in this study. A detailed analysis of diaphragm fracture under point loading may prove valuable in understanding the fracture behavior of polycrystalline diamond films. 149 VI. CLOSING REMARKS The ability to measure the mechanical properties (i.e. elastic modulus and fracture stress) of diamond films will aid in the improvement of diamond CVD processes by allowing a direct comparison of films grown under different conditions. Growing the “strongest” films is not a priority to the majority of current diamond CVD researchers, due to the relative youth of the process. As an understanding of the diamond CVD process develops and the marketplace for diamond films expands, the concerns of diamond “growers” will shift to the improvement of mechanical and other properties. Methods employed to measure the macroscopic mechanical properties of diamond films are, at this point, unable to provide reliable information. Current methods, however, have shown promise and with their improvement and the development of new techniques and equipment, the ability to measure the mechanical properties of diamond films should keep up with advancements in CVD diamond growth. The load-deflection behavior of the 17 diamond diaphragms included in this study was accurately described by the general form of existing non-linear plate theory. Although this study met with various difficulties and failed to produce an unquestionable value for the modulus of the diamond films, point loading in combination with the analysis of residual stress and buckled morphology of diamond film diaphragms could aid in determine the mechanical properties of CVD diamond and other brittle thin films. APPENDICIES APPENDIX A Equipment parameters 1. Dow Interfacial Testing System (nanoindenter) - Mitutoyo Finescope FS 100, Mitutoyo 50x objective - Sony XC-57 CCD video camera, Sony Trinitron monitor - Klinger (Garden City, NY) X-Y stepper motor - MicroControl (France) Z stepper motor - Klinger CC1 stepper motor controls. - Sartorius L—610 electronic balance - Zenith 386/20 PC - Dow interfacial testing system III software - Magwen Diamond Products (Lowell, MA) 10pm radius hemispherical diamond indenter tip. 2. Laser Scanning Confocal Microscope (LSCM) - Zeiss 1.0 LSCM (Carl Zeiss Inc., Thomwood, NY) - Single line argon ion laser - Silicon Graphics 4030 workstati0n - VoxelView graphics software (Vital Images, Fairfield, I0) 150 APPENDIX B Nanoindenter Calibration The Sartorius 610 electronic balance was tested by measuring three known masses (weights of which were verified using another electronic balance). The Sartorius 610 on the nanoindenter accurately reproduced the measured weights for the three masses. The Z stepper motor was calibrated using a glass beam specimen (Figure 46). The glass beam was cut from a microscope slide. The microscope slide, from which the calibration beam was to be cut, was chosen from a group of microscope slides used in a study by K. Lee [129] in which the elastic modulus of over 200 microscope slides was measured using a sonic resonance technique. The elastic modulus measured for the 200 microscope slides was 70 + 1 GPa. The elastic modulus and Poissons ratio of the microscope slide used for the glass beam was measured using the sonic resonance technique (K. Lee [129]) and found to be 69GPa and .18 for E and v respectively. The glass beam was cut from the microscope slide using a low speed diamond saw and the edges were ground smooth using 600 grit SiC paper. The dimensions of the glass beam (thickness = 0.95mm, width = 2.25mm) were measured using calipers and a micrometer. A thick glass bar (3 x 9 x .8 cm) was used as a base and two glass plates cut 151 152 from a coverslip were used as supports for the beam specimen. The supports were placed on the glass base and secured in place using paper/adhesive tabs. The tabs were fixed to the surface of the base plate and defined a square region in which the glass support was placed in order to prevent movement. The distance between the supports (45mm) was measured using calipers. The simply supported glass beam calibration specimen is depicted in Figure 46. The calibration specimen was placed on the stage of the nanoindenter and the center point was located. The beam was deflected using the 10pm diamond indenter tip and the corresponding load measured. After deflection of the beam, the beam was removed from the supports and placed directly on the glass base. Three indentations were made near the center of the beam in order to measure the compliance of the specimen and indenter. The measured compliance was subtracted from the load-deflection data obtained from the simply supported beam. The resulting load- deflection curve is given in Figure 47. The slope of the load deflection curve (2.5 x 10'4 m/N) for the glass beam specimen was determined using linear regression. d = deflection 3 P = measured load _d_ = __I:__ 3 L=Iength ofspan (B1) P 4Ewt E = elastic modulus of beam w = width of beam t = beam thickness Equation B.1 [128] was used to determine the predicted deflection of the beam. Equation B.1 predicts a slope of 1.7 x 10'4 m/N for the deflection/load obtained from the beam. The measured value 153 (2.5 x 10'4 m/N) for the s10pe of the deflection/load is 50% higher than the theoretically predicted value. The displacement value reported by the nanoindenter (0.04 11m per step) was multiplied by .66 to bring the measured values into agreement with beam theory. All displacement data in this study is 66% of the values reported by the nanoindenter. N S deflection (microns) 8 S B 154 0.2 0.4 0.6 0.8 1 .0 Load (Newtons) glass support \ I I P011113 103d glass beam I ./ Glass Base Figure 46. (A) Load-deflection plot for simply supported glass beam used in the nanoindenter calibration. (B) Schematic of glass beam setup. APPENDIX C Seeding of the silicon substrate. A homogenous mixture of diamond powder (0.1 - 0.2mm particles) suspended in positive photoresist (Shipley 1470J) was prepared by ultrasonic agitation. The photoresist/diamond suspension was placed on the polished silicon substrate and spun, producing a thin, even layer of photoresist and even distribution of diamond particles. After the photoresist was developed, scattered diamond particles remained on the silicon surface to act as nucleation sites for CVD diamond [127]. 155 156 APPENDIX D Grain boundary relaxation calculations for residual stress Windischmann et al. [94] used thermal expansion coefficients obtained from Slack et al. [131] to produce thermal expansion curves for silicon and diamond. The thermal expansion plots presented in the Windischmann study are similar to the plots obtained in this study using thermal expansion polynomials. Windischmann determined the room temperature thermal stress in the films by multiplying the biaxial elastic modulus of diamond (1345GPa) [94] by the predicted thermal mismatch between diamond and silicon at the diamond deposition temperature. The values of thermal strain used by Windischmann et al. do not agree with the thermal expansion plots produced in the Windishmann study or this study and are likely the result of a miscalculation. The predicted thermal stress values in the Windischmann study when replaced by those of Toloukian [4], depicted in Figure 10, produce the following results: Gin = intrinsic stress Gin = O'T - 6th CT = total stress (D.1) 0th = thermal stress The total stress at room temperature was determined through measurement of the curvature induced in the silicon wafer by the 157 film combined with the Stoney equation and is believed to be accurate to 10% [94]. For a film with an. average grain. size of approximately 0.12pm grown at ~790°C the total residual stress was 0.2GPa. Figure 10 predicts a thermal stress (thermal strain x biaxial elastic modulus for diamond) of (5.8 x 10'4 x 1345 GPa) = .78 GPa. The value of .7 8 GPa is within 10% of the value (.84 GPa) predicted by the grain boundary relaxation model (Equation 2.1). LIST OF REFERENCES LIST OF REFERENCES P. D. Ownby, “Engineering Properties of Diamnd and Graphite,” ASM Engineered Materials Handbook Volume 4, (ASM International, 1991), Pp. 821-834. J. E. Field, “Natural Diamond: The Standard,” Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991), PP. 17-35. M. D. Drory, C. F. Gardinier, and J. M. Pinneo, “Fracture Behavior of CVD Diamond,” Thin Films: Stresses and Mechanical Properties III, Materials Research Society Symposium Proceedings Volume 239, edited by W. D. Nix, J. C. Bravman, E. Arzt, and L. B. Freund, (Materials Research Society, 1992), pp. 561-566. Y. S. Touloukian, R. E. Kirby, R. E. Taylor, and T. Y. R. Lee, Thermophysical Properties of Matter; TPRC Series, (IFI/Plenum 1977 ). N. K. Annamalai, R. F. Blanchard, and Joseph Chapski, “Radiation Response of Diamond Films,” Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991), pp. 821-828. C. P. Beetz, C. V. Cooper, and T. A. Perry, “Ultralow Load Indentation Hardness and Modulus of Diamond Films Deposited by Hot-Filament- Assisted CVD,” Journal of Materials Research, 5: 2555-2561, (1990). P. K. Bachmann and R. Messier,”Emerging Technology of Diamond Thin Films,” Chemical and Engineering News, 70 (20):24-39, (1989). M. N. Yoder, “Diamond: Potential and Status,” Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991), pp. 1-16. K. Nishimura, K. Das, and J. T. Glass, “Material and Electrical Characterization of Polycrystalline Boron-Doped Diamond Films Grown By Microwave Plasma Chemical Vapor Deposition,” Journal of Applied Physics, 69: 3142-3148, (1991). 158 10. 11. 12. 13. 14. 15. 16. 17. -18. 19. 20. 159 R. N. Boggs, “Diamond Thin Film: Hot New Material For The 90’s,” Design News, 45 (7): 70-75, (1989). R. Pool, “Diamond Films Sparkle As They Come to Market,” Science, 249: 27-28, (1990). A. E. Alexenko and B. V. Spitsyn, “Some Properties of CVD Diamond Semiconducting Structures,” Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press,1991), pp.789-798. H, Kiyota, K. Okano, T. Iwasaki, H. Izumiya, Y. Akiba, T. Kurosu, and M. Iida, “Fabrication of Metal-Insulator-Semiconductor Devices Using Polycrystalline Diamond Film,” Japanese Journal of Applied Physics, 30: L2015-L2017, (1991). K. Okano, H. Kiyota, T. Iwasaki, T. Kurosu, M Iida, and T. Nakamura, “p-n Junction Diode Made of Semiconducting Diamond Films,” Applied Physics Letters, 58: 840-841, (1990). R. C. McCune, D. W. Hoffman, T. J. Whalen, and C. O. McHugh, “Adherence of Diamond Films Produced by Microwave Plasma Deposition on SiAlON Tool Inserts,” Thin Films: Stresses and Mechanical Properties, Materials Research Society Symposium Proceedings Volume 130, edited by J. C. Bravman, W. D. Nix, D. M. Barnett, and D. A. Smith, (Materials Research Society, 1989),pp. 261-266. P. W. Bridgman, “Synthetic Diamonds,” Scientific American, 193: 42- 46, (1955). ’ F. P. Bundy, H. T. Hall, H. M. Strong, and R. J. Wentorf, “Manmade Diamond,” Nature, 176: 51-55, (1955). W. Zhu, B. R. Stoner, B. E. Williams, and J. T. Glass, “Growth and Characterization of Diamond Films on Nondiamond Substrates for Electronic Applications,” Proceedings of the IEEE: 79: 621-646, ( 1991). J. C. Angus and C. C. Hayman, “Low-Pressure, Metastable Growth of Diamond and Diamondlike Phases,” Science, 241: 913-921, (1988). K. E. Spear, “Diamond- Ceramic Coating of the Future,” Journal of the American Ceramic Society, 72: 171-191, (1989). 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 160 T. R. Anthony, “Methods of Diamond Making,” Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991), pp. 555-577. W. G. Eversole, “Synthesis of Diamond by Deposition on Seed Crystals,” U. S. Patent #3030187, (1962). W. A. Yarbrough and R. Messier, “Current Issues and Problems in the Chemical Vapor Deposition of Diamond,” Science, 247 :688-695, (1990). J. C. Angus, H. A. Will, and W. S. Stanko, “Growth of Diamond Seed Crystals by Vapor Deposition,” Journal of Applied Physics, 3922915- 2922, (1968). N. Setaka, “Critical Assessment of State-of-the-Art of Growing Diamond,” Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991), pp. 875-888. B. V. Spitsyn, L. L. Bouilov, and B. V. Derjaguin, “Vapor Growth of Diamond on Diamond and Other Surfaces,” Journal of Crystal Growth, 52: 219-226, (1981). P. K. Bachmann and H. Lydtin, “High Rate Versus Low Rate Diamond CVD Methods,” Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991), pp. 829-854. B. V. Spitsyn, “Origin and Evolution of the Science and Technology of Diamond Synthesis in the USSR,” Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991). pp. 855-873. A. R. Badzian and R. C. DeVries, “Crystallization of Diamond From the Gas Phase; Part 1,” Materials Research Bulletin, 23: 385-400, (1988). J. Mercier, A. M. Bonnot, E. Caignol, and E. Gheeraert, “Investigations of Growth Mechanisms of Diamond Thin Films by Hot Filament Assisted CVD,” Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991), pp. 533- 540. 31. 32. 33. 34. 35. 36. 37. 38. 39. 161 S. J. Harris, A. M. Weiner, and T. A. Perry,“ Filament-Assisted Diamond Growth Kinetics,” Journal of Applied Physics, 70: 1385- 1391, (1991). K. Baba, Y. Aikawa, and N. Shohata, “Thermal Conductivity of Dia- mond Films,” Journal of Applied Physics, 69: 7313-7315, (1991). P. K. Bachamnn and D. U. Wechert, “Characterization and Properties of Artificially Grown Diamond,” Diamond and Diamond-Like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991). pp. 677-714. P. Bou and L. Vandenbulcke, “Raman Investigations on Diamond Films and Crystals Deposited by Plasma-Assisted CVD,” Journal of the Electrochemical Society, 138: 2991-3000, (1991). F. G. Celii, D. White, and A. J. Purdes, “Effect of Residence Time on Microwave Plasma Chemical Vapor Deposition of Diamond,” Journal oprplied Physics, 70: 5636-5646, (1991). Y. Saito, S. Matsuda, and S. N ogita, “Synthesis of Diamond by Decomposition of Methane in Microwave Plasma,” Journal of Materials Science Letters, 5: 565-568, (1986). Y. Liou, A. Inspektor, R. Weimer, D. Knight, and R. Messier, “The Effect of Oxygen in Diamond Deposition by Microwave Plasma Enhanced Chemical Vapor Deposition,” Journal of Materials Research, 5: 2305-2312, (1990). T. Mitomo, T. Ohta, E. Kondoh, and K. ohtsuka, “An Investigation of Product Distributions in Microwave Plasma for Diamond Growth,” Journal of Applied Physics, 70: 4532-4539, (1991). M. Nunotanim M. Komori, M. Yamasawa, Y. Fujiwara, K. Sakuta, T. Kobayashi, S. Nakashima, S. Minomo, M. Taniguchi, and M. Sugiyo, “Effects of Oxygen Addition on Diamond Film Growth by Electron- Clyclotron-Resonance Microwave Plasma CVD Apparatus,” Japanese Journal of Applied Physics, 30: L1199-L1202, (1991). J. J. Chang, T. D. Mantei, R. Vuppuladhadium, and H. E. Johnson, “Low Temperature and Low Pressure Diamond Systhesis in a Microwave Electron Cyclotron Resonance Discharge,” Applied Physics Letters, 59: 1170-1172, (1991). 41. 42. 43. 45. 46. 47. 48. 49. 50. 51. 162 H. Kawarada, K. S. Mar, and A Hiraki, “Large Area Deposition of Diamond Particles and Films Using Magneto-Microwave Plasma,” Japanese Journal of Applied Physics, 26: L1032-L 1034, (1987). M. Kawarada, K. Kurihara, and K. Sasaki, “Synthesis of a Diamond and Metal Mixture by the Chemical Vapor Deposition Process,” Journal of Applied Physics, 71: 1442-1445, (1992). F. Akatsuka, Y. Hirose, and K. Komaki, “Rapid Growth of Diamond Films by Arc Discharge Plasma CVD,” Japanese Journal of Applied Physics, 27: L1600-L1602, (1988). K. Kurihara, K. Sasaki, M. Kawarada, N. Koshino, “High Rate Synthesis of Diamond by DC Plasma Jet Chemical Vapor Depositon,” Applied Physics Letters, 52: 437-438, (1988). Y. Hirose, S. Amanuma, and K. Komaki, “The Synthesis of High- Quality Diamond in Combustion Flames,” Journal of Applied Physics, 68: 6401-6405, (1990). K. V. Ravi, C. A. Koch, H. S. Hu, and A. J oshi, “The Nucleation and Morphology of Diamond Crystals and Films Synthesized by the Combusiton Flame Technique,” Journal of Materials Research, 5: 2356-2366, (1990). M. Frenklach, “The Role of Hydrogen in Vapor Deposition of Diamond,” Journal of Applied Physics, 65: 5142-5149, ( 1989). M. Frenklach, “Theory and Models for Nucleation and Growth of Diamond Films,” Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991). Pp. 499-524. Z. Chen, G. P. Wirtz, and S. D. Brown, “Chemical Kinetic Calculation of the Gas Phase in Filament-Assisted Chemical Vapor Deposition of Diamond,” Journal of the American Ceramic Society, 75: 2107-2115, (1992). . C. J. Chu, M. P. D’Evelyn, R. H. Hauge, and J. L. Margrave, “Mechanism of Diamond Growth by Chemical Vapor Depositon on Diamond (110), (111), and (110) surfaces: Carbon-13 Studies,” Journal of Applied Physics, 70: 1695-1705, (1991). L. R. Martin, “Diamond Film Growth in a Flowtube: A Temperature Dependance Study,” Journal of Applied Physics, 70: 5667-5674, (1991). 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 163 M. Frenklach and K. E. Spear, “Growth Mechanism of Vapor- Deposited Diamond,” Journal of Materials Research, 3: 133-140, (1988). ‘ L. Vandenbulcke, P. Bou, and G. Moreau, “On the Control of Diamond Deposition Structure and Morphology,” Journal of the Electrochemical Society, 139: 2985-2991, (1991). S. J. Harris, “Gas-Phase Kinetics During Diamond Growth: CH4 As- Growth Species,” Journal of Applied Physics, 65: 3044-3048, (1989). S. J. Harris, L. R. Martin, “Methys Versus Acetylene as Diamond Growth Species,” Journal of Materials Research, 5: 2313-2319, (1990). L. R. Martin, M. W. Hill, “A Flow-Tube Study of Diamond Film Growth: Methane Versus Acetylene,” Journal of Materials Science Letters, 9: 621-623, (1990). D. G. Goodwin and G. G. Gavillet, “Numerical Modeling of the Filament-Assisted Diamond Growth Environment,” Journal of Applied Physics, 68: 6393-6400, (1990). T. Kawato and K. Kondo, “Effects of Oxygen on CVD Diamond Synthesis,” Japanese Journal of Applied Physics, 26: 1429-1432, (1987 ). J. A. Mucha, D. L. Flamm, and D. E. Ibbotson, “On The Role of Oxygen and Hydrogen in Diamond-Forming Discharges,” Journal of Applied Physics, 65: 3448—3452, (1989). B. Lux and R. Haubner, “Nucleation and Growth of Low Pressure Diamond,” Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991), pp. 57 9-609. M. Frenklach, R. Kamatick, D. Huang, W. Howard, K. E. Spear, A. W. Phelps, and R. Koba, “Homogeneous Nucleation of Diamond Powder in the Gas Phase,” Journal of Applied Physics, 66: 395-398, (1989). C. P. Chang, D. L. Flamm, D. E. Ibbotson, and J. A. Mucha, “Diamond Crystal Growth by Plasma Chemical Vapor Deposition,” Journal of Applied Physics, 63: 1744-1748, ( 1988). 63. 64. 66. 67. 68. 69. 70. 71. 72. 73. 74. 164 Y. Muranaka, H. Yamashita, and H. Miyadera, “Characterization of Diamond Films Synthesized in the Microwave Plasmas of CO/H2 and CO/Oz/Hg Systems at Low Temperatures (403-1023 K),” Journal of Applied Physics, 68: 8145-8153, (1991). ' Y. Saito, K. Sato, H. Tanaka, K. Fujita, and S. Matuda, “Diamond Synthesis from Methand-Hydrogen-Water Mixed Gas Using a Microwave Plasma,” Journal of Materials Science Letters, 23: 842-846, (1988). C. J. Tomg, J. M. Sivertsen, J. H. Judy, and C. Chang, “Structure and Bonding Studies of the C:N Thin Films Produced by RF SPuttering Method,” Journal of Materials Research, 5: 2490-2496, (1990). R. Meilunas, R. Chang, S. Liu, and M. Kappes, “Nucleation Control of Diamond Films on Surfaces Using Carbon Clusters,” Applied Physics Letters, 59: 3461-3463, (1991). H. Itoh, T. Osaki, H. Iwahara, and S. Sakamoto, “Nucleation Control of Diamond Synthesized by Microwave Plasma CVD on Cemented Carbide Substrate,” Journal of Materials Science, 26: 3763-37 68, (1991). D. J. Pickrell, W. Zhu, A. R. Badzian, R. E. N ewnham, and R. Messier, “Near Interface Characterization of Diamond Films on Silica and Silicon,” Journal of Materials Research, 6: 1264-127 7, (1991). S. Iijima, Y. Aikawa, and K. Baba, “Growth of Diamond Particles in Chemical Vapor Deposition,” Journal of Materials Research, 6: 1491- 1497, (1991). K. Hirabayashi, Y. Taniguchi, O. Takamatsu, T. Ikeda, K Ikoma, and N. Iwasaki-Kurihara, “Selective Depositon of Diamond Crystals by Chemical Vapor Depositon Using a Tungsten-Filament Method,” Applied Physics Letters, 53: 1815-1817, (1988). S. Matsumoto, Y. Matsui, “Electron Microscopic Observation of Diamond Particles Grown From the Vapour Phase,” Journal of Materials Science, 18: 1785-1793, (1983). M. A. Tamor and K. C. Hass, “Hypothetical Superhard Carbon Metal,” Journal of Materials Research, 5: 2273-227 6, (1990). B. E. Williams and J. T. Glass, “Characterization of Diamond Thin Films: Diamond Phase Identification, Surface Morphology, and Defect Structures,” Journal of Materials Research, 4: 373-384, (1989). 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 165 J. Narayan, A. R. Srivatsa, M. Peters, S. Yokota, and K. V. Ravi, “On Epitaxial Growth of Diamond Films on (100) Silicon Substrates,” Applied Physics Letters, 53: 1823-1825, (1988). D. G. K. Jeng, H. S. Than, J. E. Butler, R. F. Salat, and G. J. Fricano, “Local Heteroepitaxial Diamond Growth on (100) Silicon,” Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991), pp. 627-634. I A. R. Badzian, “Defect Structure of Synthetic Diamond and Related Phases,” Advances in X-Ray Analysis, Volume 31, edited by C. B. Barrett, J. V. Gilfrich, R. Jenkins, J. W. Richardson, and P. K. Predecki, (Pleneum Press, 1988), pp. 113-128. D. E. Peebles and L .E. Pope, “Analytical and Mechanical Evaluation of Diamond Films on Silicon,” Journal of Materials Research. 5: 2589- 2598, (1990). C. T. Kuo, T. Y. Yen, T. H. Huang, and S. E. Hsu, “Adhesion and 'I‘ribological Properties of Diamond Films on Various Substrates,” Journal of Materials Research, 5: 2515-2523, (1990). D. W. Kweon and J. Y. Lee, “The Interface Morphology of Diamond Film Grown on Si,” Journal of Applied Physics, 68:4272-4275, (1990). K Kobashi, K. Nishimura, Y. Kawate, and T. Horiuchi, “Synthesis of Diamonds By the Use of Microwave Plasma Chemical-Vapor Deposition; Morphology and Growth of Diamond Films,” Physical Review B, 38: 4067—4084, (1988). ‘ C. Wild, N. Herres, and P. Koidl, “Texture Formation in Polycrystalline Diamond Films,” Journal of Applied Physics, 68: 973- 978, (1990). C. Tsai, J. Nelson, W. Gerberich, J. Heberlein, and E. Pfender, “Metal Reinforced Diamond Composite Films,” Thin Films: Stresses and Mechanical Properties III, Materials Research Society Symposium Pro- ceedings Volume 239, edited by W. D. Nix, J. C. Bravman, E. Arzt, and L. B. Freund, (Materials Research Society, 1992), pp. 275-280. J. Narayan, “Dislocations, Twins, and Grain Boundaries in CVD Diamond Thin Films: Atomic Structure and Properties,” Journal of Materials Research, 5: 2414-2423, (1990). 85. 86. 87. 88. 89. 90. 91. 92. 93. 166 E. D. Specht, R. E. Clausing, and L. Heatherly, “Measurement of Crystalline Strain and Orientatin in Diamond Films Grown By Chemical Vapor Deposition,” Journal of Materials Research, 5: 2351- 2355, (1990). Y. Ohsawa, Y. Tamou, N. Kikuchi, K. Hiraga, T. Oku, “TEM Observa- tions of Diamond Films Produced by Hot Filament Thermal CVD,” Journal ofMaterials Science, 26: 3748-3752, (1991). S. J. Harris, D. N. Belton, R. J. Blint, “Thermochemistry on the Hydrogenated Diamond (111) Surface,” Journal of Applied Physics, 7 0: 2654-2659, (1991). H. Hirai, O. Fukunaga, and O. Odawara, “Effect of Microwave Power on Hydrogen Content in Chemically Vapor Deposited Diamond Films,” Journal of the American Ceramic Society, 74: 1715-1718, (1991). D. S. Knight and W. B. White, “Characterization of Diamond Films by Raman Spectroscopy,” Journal of Materials Research, 4: 385-393, (1989). T. Anthony, W. Banholzer, J. Fleischer, L. Wei, P. Kuo, R. Thomas, and R. Pryor, “Thermal Diffusivity of Isotopically Enriched 12C Diamond,” Physical Review B, 42: 1104-1111, (1990). K. M. McNamara and K. G. Gleason, “Carbon Bonding Environments in CVD Diamond Films Investigated Via Solid-State NMR”, Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991), pp. 541-547. G. J. Lewinsink, T. G. M. Oosterlaken, G. C. A. M. Janssen, and S. Radelaar, “Evolution of Intrinsic Stress During Nucleation and Growth of Polycrystalline Tungsten Films by Chemical Vapor Deposition,” Thin Films: Stresses and Mechanical Properties III, Materials Research Society Symposium Proceedings Volume 239, edited by W. D. Nix, J. C. Bravman, E. Arzt, and L. B. Freund, (Materials Research Society, 1992), pp. 139-144. R. S. Fisher, J. Z. Duan, and A. G. Fox, “Stress Gradients and Anisotropy in Thin Films,” Thin Films: Stresses and Mechanical Properties, Materials Research Society Symposium Proceedings Volume 130, edited by J. C. Bravman, W. D. Nix, D. M. Barnett, and D. A.'Smith, (Materials Research Society, 1989), pp. 249-254. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 167 H. Windischmann, G. F. Epps, Y. Cong, and R. W. Collins, “Intrinsic Stress in Diamond Films Prepared by Microwave Plasma CVD,” Journal of Applied Physics, 69: 2231-2237, (1991). B. S. Berry, W. C. Pritchet, J. J. Cuomo, C. R. Guarnieri, and S. J. Whitehair, “Intrinsic Stress and Elasticity of Synthetic Diamond Films,” Applied Physics Letters, 57: 302-303, (1990). A. S. Argon, V. Gupta, H. S. Landis, J. A. Cornie, “Intrinsic Toughness of Interfaces Between SiC Coatings and Substrates of Si or C Fibre,” Journal ofMaterials Science, 24: 1207-1218, (1988). S. Hong, T. P. Wheis, J. C. Bravman, and W. D. Nix, “Measuring Stiffnesses and Residual Stresses of Silicon Nitride Thin Films,” Journal of Electronic Materials, 19: 903-909, (1990). H. Windischmann, “Intrinsic Stress in Sputter Deposited Thin Films,” Critical Reviews in Solid State and Materials Science, 17: 547-592, (1992). J. L. Davidson, R. Ramesham, and C. Ellis, “Synthetic Diamond Micromechanical Membranes, Cantilever Beams, and Bridges,” Journal of the Electrochemical Society, 137: 3206-3210, (1990). G. F. Cardinale and R. W. Tustison, “Biaxial Modulus Measurement of Chemical Vapor Deposited Polycrystalline Diamond Films,” Journal of Vacuum Science and Technology A, 9: 2204-2209, (1991). CT. Rosenmayer, F. R. Brotzen, and R. J. Gale, “Mechanical Testing of Thin Films,” Thin Films: Stresses and Mechanical Properties, Materials Research Society Symposium Proceedings Volume 130, edited by J. C. Bravman, W. D. Nix, D. M. Barnett, and D. A. Smith, (Materials Research Society, 1989), DP. 7 7-86. D. S. Stone, K. B. Yoder, and W. D. Sproul, “Hardness and Elastic Modulus of TiN Based on Continuous Indentation Technique and New Correlation,” Journal of Vacuum Science and Thchnology A, 9: 2543- 2547, (1991). D. L. J oslin and W. C. Oliver, “A New Method for Analyzing Data From Continuous Depth-Sensing Microindentation Tests,” Journal of Materials Research, 5: 123-126, (1990). W. W. Weiler, “Dynamic Loading: A New Microhardness Test Method,” Journal of Testing and Evaluation, 18 (4): 229-239, (1990). 105. 106. 107. 108. 109. 110. 111. 113. 168 M. E. O’Hern and C. J. McHargue, “Mechanical Properties Testing of Diamond and Diamond-like Films by Ultra-Low Load Indentation,” Diamond and Diamond-like Films and Coatings, NATO ASI Series B. Volume 266, edited by R. E. Clausing, L. L. Horton, J. C. Angus, and P. Koidl, (Plenum Press, 1991), pp. 715-722. B. S. Berry and W. C. Pritchet, “Internal Stress and Internal Friction in Thin-Layer Microelectronic Materials,” Journal of Applied Physics, 67: 3661-3668, (1990). X. Jiang, J. V. Harzer, B. Hillebrands, Ch. Wild, and P. Koidl, “Brillouin Light Scattering on Chemical-Vapor-Deposited Polycrystalline Diamond, Evaluation of the Elastic Moduli,” Applied Physics Letters, 59: 1055-1057, (1991). X. Jiang, “Brillouin Light Scattering on Diamond and Diamondlike Carbon Films,” Thin Films: Stresses and Mechanical Properties III, Materials Research Society Symposium Proceedings Volume 239, edited by W. D. Nix, J. C. Bravman, E. Arzt, and L. B. Freund, (Materials Research Society, 1992), pp. 269-274. T. P. Weihs, S. Hong, J. C. Bravman, and W. D. Nix, “Measuring the Strength and Stiffness of Thin Film Materials by Mechanically Deflecting Cantilever Microbeams,” Thin Films: Stresses and Mechanical Properties, Materials Research Society Symposium Proceedings Volume 130, edited by J. C. Bravman, W. D. Nix, D. M. Barnett, and D. A. Smith, (Materials Research Society,1989), PP.87-92. S. Hong, T. P. Weihs, J. C. Bravman, and W. D. Nix, “The Determination of Mechanical Parameters and Residual Stresses for Thin Films Using Micro-Cantilever Beams,” Thin Films: Stresses and Mechanical Properties, Materials Research Society Symposium Proceedings Volume 130, edited by J. C. Bravman, W. D. Nix, D. M. Barnett, and D. A. Smith, (Materials Research Society, 1989), PP.93-98. M. K. Small, J. J. Vlassak, and W. D. Nix, “Re-Examining the Bulge Test: Methods for Improving Accuracy and Reliability,” Thin Films: Stresses and Mechanical Properties III, Materials Research Society Symposium Proceedings Volume 239, edited by W. D. Nix, J. C. Bravman, E. Arzt, and L. B. Freund, (Materials Research Society, 1992), PP. 257-262. S. Timoshenko and S. Woinowsky-Krieger, Theory of Plates and Shells, (McGraw Hill, 1959). 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 169 H. Reismann, Elastic Plates: Theory and Application, (John Wiley & Sons, 1988). S. Way, “Bending of Circular Plates With Large Deflection,” Transactions of the American Society of Mechanical Engineers, 56: 627-636, (1934). C. W. Bert and J. L. Martindale, “An Accurate, Simplified Method for Analyzing Thin Plates Undergoing Large Deflections,” AIAA Journal, 26 (2): 235-241, (1988). B. Banerjee, “Large Deflection of a Circular Plate Under A Concentrated Load - A New Approach,” Journal of the Industrial Mathematics Society, 33(1): 57-61, ( 1983). R. Schmidt, “Large Deflections of a Clamped Circular Plate,” Journal of Engineering Mechanics Division, Proceedings of The AS CE, 94: 1603-1606, (1968). C. Y. Chia, Nonlinear Analysis of Plates, (McGraw Hill, 1980). B. Banerjee and S. Datta, “A New Approach to an Analysis of Large Deflections of Thin Elastic Plates,” International Journal of Non- Linear Mechanics, 16(1): 47-51, (1981). B. Banerjee, “Large Deflections of Circular Plates of Variable Thickness,” International Journal of Solid Structures, 19(2): 179-182, (1983). J. Mattavi, “Modeling of the Strain of Diamond Thin-Films on Silicon Substrates,” Senior Research and Design Project, Depatement of Materials Science and Mechanics, Michigan State University, (1993). G. M. Pharr, “The Anomalous Behavior of Silicon During N anoindentation,” Thin Films: Stresses and Mechanical Properties III, Materials Research Society Symposium Proceedings Volume 239, edited by W. D. Nix, J. C. Bravman, E. Arzt, and L. B. Freund, (Materials Research Society, 1992), pp. 301-312. S. Timoshenko and J. Gere, Theory of Elastic Stability, (McGraw-Hill, 1961), pp. 390. L. E. D. Flowers, Graduate Student, Michigan State University, Department of Materials Science and Mechanics, Personal Communication, (1992). 126. 127. 128. 129. 130. 131. 170 J. Zhang, Graduate Student, Michigan State University, Department of Electrical Engineering, Personal Communication (1993‘). A. Masood, M. Aslam, M. Tamor, and T. Potter, “Techniques for Patterning of CVD Diamond Films on Non-Diamond Substrates,” Journal of the Electrochemical Society, 138: L67-L71, (1991). Y. Ku, Deflections of Beams for all Spans and Cross Sections, (McGraw- Hill, 1986), pp. 17. K Lee, “Dependance of the Effective Young’s Modulus of Glass/Epoxy and Glass/Glue Laminants on Adhesion Area and Glue Bond Thickness”, Masters Thesis, Department of Materials Science and Mechanics, Michi- gan State University, 1993. R. Averill, Professor of Materials Science and Mechanics, Michigan State University, Personal Communication, (1993). G. A. Slack and S. F. Bartram, “Thermal Expansion of Some Diamondlike Crystals”, Journal of Applied Physics, 46:89-98, 1975. "Illllllllllllllllllll“