$31... . www.mwwvfidmgm ... sfi‘d‘ 4. . an” :uw .. .5, a. .1... ., . . . . . . , ‘ . . . . , ‘ . . x A ‘ :7. €AY§2§§Q 4.‘ 7 §§ . a 3 .u a 11¢ .. flag... . , s s :3 . ‘ . a. L . . . 1:“. 3 . ‘ avvéifi .,;U.WF..-.. .3 ‘ 5?. A. -) $3. .flerraf dwfimmau 4......Vfl ma. v t .x. .J .t a. t , .2 a twitrmpiu . gnuflfg..$r%.§kflwln A” if... 1.. ‘ .V . , . m5». 2.- .. 1.2.. This is to certify that the thesis entitled The Influence of the Microstructure of Sintered Hydroxyapatite on the Properties of Hardness, Fracture Toughness, Thermal Expansion and the Dielectric Permittivity presented by TImothy P. Hoepfner has been accepted towards fulfillment of the requirements for Master's degree in Materials Science and Engineering W09: W Ward as, 200/ 0-7639 MS U is an Affirmative Action/Equal Opportunity Institution LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE AWi f 2123‘. F?12 1.2.0 9 601 CJCIRC/DataDuopes-ptis THE INFLUENCE OF THE MICROSTRUCTURE OF SINTERED HYDROXYAPA'ITTE ON THE PROPERTIES OF HARDNESS, FRACTURE TOUGHNESS, THERMAL EXPANSION AND THE DIELECTRIC PERMITTIVITY By Timothy P. Hoepfner 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 April 2001 ABSTRACT THE INFLUENCE OF THE MICROSTRUCTURE OF SINTERED HYDROXYAPATITE ON THE PROPERTIES OF HARDNESS, FRACTURE TOUGHNESS, THERMAL EXPANSION AND THE DIELECTRIC PERMITTIVITY By Timothy P. Hoepfner Sintered hydroxyapatite specimens with a 58-98% relative density range and average grain sizes between 1.64 um and 7.44 pm were made with a commercial hydroxyapatite powder. No decomposition of the hydroxyapatite was detected, by x-ray difli'action, for any specimen. Vickers hardness (Hv) of the Sintered hydroxyapatite specimens was successfully fit, as a ftmction of the volume fi'action porosity (P), to the minimum solid area model (MSA): Hv = HVO 6"". The zero porosity hardness value Hv0 = 6.00 :i: 0.79 GPa and the material dependent constant b = 6.03. Fracture toughness (ch0) of the Sintered hydroxyapatite specimens, measured by indentation methods, diverged fi'om the MSA model. Fracture toughness increased as the porosity increased. Kw0 a 0.5 MPa m"2 for Sintered hydroxyapatite specimens of 97% relative density increasing to K“,0 22.6 MPa In"2 for specimens of 72% relative density. Thermal expansion coefficients for Sintered hydroxyapatite, measured by dilatometry = 12.9 x 10" / °C at 100° C , 13.8 x 1045 / °C at 200° C, 16.3 at 400° C and 19.8 x 1045 / °C at 600° C. The relative dielectric permittivity was measured for several Sintered hydroxyapatite specimens of varying porosity and the zero porosity relative dielectric permittivity was calculated using a dielectric mixing law to k0 = 17.63 :t 0.70. TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES 1.0 Introduction 1.1 1.2 1.3 1.4 1.5 Goals of research Chemistry of sintering 1.2.1 Dehydroxylation and decomposition 1.2.2 Process control of phase purity during sintering Relevant crystal structures 1 .3.1 Hydroxyapatite 1.3.2 Tricalcium phosphate 1.3.3 Tetracalcium phosphate Microstructure of Sintered hydroxyapatite 1.4.1 Residual strain energy and microcracking Microstructure and material Properties 1.5.1 Indentation hardness 1.5.2 Indentation method to determine fi'acture toughness 1.5.3 Hardness and fi'acture toughness studies 1.5.4 Fracture energy and microstructure 1.5.5 Indentation studies of Sintered hydroxyapatite 1.5.6 Summary of hardness and toughness studies 1.5.6.1 Hardness 10 10 10 13 16 17 21 21 2.0 1.6 1.7 1.8 1.5.6.2 Fracture toughness Material properties and porosity and the minimum solid area Thermal expansion and microstructure 1.7.1 Thermal expansion and microcracking Dielectric properties and micro structure Experimental Procedure 2.1 2.2 2.3 2.4 Starting material 2.1.1 Particle size analysis 2.1.2 Ball milling Specimen preparation 2.2.1 Die forming 2.2.2 Cold isostatic pressure 2.2.3 Conventional filmace sintering 2.2.4 Microwave sintering 2.2.5 Sectioning the Sintered specimens 2.2.6 Specimen polishing 2.2.7 Surface etching 2.2.8 Specimen mounting for scanning electron microscopy Microstructure measurements 2.3.1 Mass density 2.3.2 Grain size measurements 2.3.3 X-ray difli'action Hardness and fi‘acture toughness measurements iv 21 22 24 25 25 27 27 27 28 28 30 30 37 39 41 43 45 45 45 46 48 50 3.0 2.5 2.6 Thermal expansion 2.5.1 Specimen preparation for thermal expansion measurements 2.5.2 Thermal expansion measurement Dielectric measurements 2.6.1 Specimen preparation for dielectric measurements 2.6.2 Dielectric apparatus and measurements Results and Discussion 3.1 3.2 3.3 3.4 3.5 Sintered hydroxyapatite micro structure 3.1.1 Sintered density 3.1.2 Sintered grain size 3.1.3 Etched surface morphology 3.1.4 X-ray diffraction Indentation measurements 3.2.1 Vickers hardness 3.2.2 Indentation fracture toughness 3.2.2 Thermal expansion Hardness and microstructure comparisons with literature data Fracture toughness and microstructure comparisons with literature data Further comment on the fracture toughness of Sintered hydroxyapatite 3.5.1 Indentation toughness model and porosity 3.5.1.1 Efl‘ect of E(Pp) on the fracture toughness of Sintered hydroxyapatite 51 51 51 52 52 52 56 56 56 56 76 76 81 81 91 98 99 104 3.5.1.2 Indentation fracture toughness of porous YBa.2Ca3O7_x 3.5.1.3 Porosity dependence of the empirical constant C 3.5.2 Examples of deviations fiom the MSA model for fiacture toughness 3.5.3 Possible toughening mechanisms in Sintered hydroxyapatite 3.6 Thermal expansion 3 .7 Dielectric measurements 3.7.1 Circuit capacitance measurement and correction 3.7.2 The relative permittivity of hydroxyapatite 4.0 Summary and Conclusions 4.1 Further research Appendix A Appendix B References 106 110 114 114 120 128 128 129 135 137 139 147 154 Table 1-1 Table 1-2 Table 2-1 Table 2-2 Table 2-3 Table 2-4 Table 2-5 Table 3-1 Table 3-2 LIST OF TABLES A survey of secondary phase formation in Sintered hydroxyapatite under varied sintering conditions for various studies from the literature. Ceramic oxides in which the bulk thermal expansion coefficient was constant for a variety of grain sizes and porosity. Specific analysis (Ca percentages) and spectro graphic analysis (impurities) of the hydroxyapatite powder recieved fi'om CERAC Inc., Milwaukee, WI. The data is copied from the Certificate of Analysis attached to material. Processing parameters for Hydroxyapatite disk specimens. Fired in automatically controlled tube furnace. HAP I is lot number X16907 from CERAC inc. Milwaukee, WI. Processing parameters for rectangular HAP Specimens, fired in C+M box furnace under manual control. HAP II is from lot number X21949 fiom CERAC. HAP Ia, HAP IIa- hydroxyapatite powder ball milled with alumina pellets, HAP Iz- hydroxyapatite powder ball milled with zirconia pellets. Processing parameters for disk Specimens, fired in C+M box furnace under manual control Processing parameters for disk HAP Specimens fired in a microwave cavity. Processing parameters, density and grain size data for Sintered hydroxyapatite. Lattice parameter data for the as received hydroxyapatite powder (HAP I) and Sintered hydroxyapatite, calculated using the 3 major hydroxyapatite peaks fiom Figures 3-13 and 3-14 and Equations 2-2 and 2-3. Data fiom the handbook is included [69]. 24 27 31 32 33 34 57 Table 3-3 Table 3-4 Table 3-5 Table 3-6 Table 3-7 Table 3-8 Table 3-9 Density, grain size and the mechanical properties of hardness and toughness, measured by indentation technique, of Sintered hydroxyapatite. The standard deviations (0), calculated for the multiple indentations per specimen, are included. 78 The normalization and statistical parameters fiom the linear regression analysis of the hardness as a function of volume fi'action porosity (Equation 3-1: In H = 1n Ho - bP) for this study, Wang et al. [9], Slosarczyk et al. [11], Best et al. [75], and Lu et al. [74]. Actual hardness data [9,11,74,75] from the literature is included. 88 The normalization and statistical parameters fi'om the linear regression analysis of toughness as a function of volume fraction porosity (Equation 3-2: In K,c = 1n K0ic - b P) for this study, DeWith et al. [54] and Suchanek et al. [56]. Fracture toughness data from DeWith et al. [54] and Suchanek et al. [56] are included. 92 Porosity dependence of the elastic modulus compared to the porosity dependence of Poisson’s ratio for experimental data fit to the MSA model (lnE = in E0 - bEP, lnv = in v0 - va, Equation 1-8) by Rice [58]. 101 Material dependent coefficients (b,.l and b5) from the MSA model [57] for A1203, B4C, AIN and SiO2 ceramics calculated by Rice [58]. The HAP (hydroxyapatite) elastic modulus attenuation coefficient was calculated for this study using DeWith et al.’s data [54]. The hardness attenuation coefficient for HAP has been determined previously in this study (Table 3-4). 103 Experimental data for YBa¢Cu3O7,x from Tancret et al. [78] and the fit, for this study, to the MSA model ln(A) = ln(Ao)- bAPP, where A represents the property (E, Hv, etc.), A0 = the property at zero porosity (y-intercept for the linear regression), bA is the slope of the log-linear relation, and Pp is the volume fiaction porosity. Included for comparison are the fracture toughness values calculated, for each porosity, using equation 3-3 {ch = C(E/H)"2 P cm}, for this study. Column 6 contains the fiacture toughness calculated with constant elastic modulus (E =120 GPa), {K‘candent constant E)}. Fracture toughness values in column 7 were calculated using the measured specimen elastic modulus [78], in equation 3-3 {K,c(lndent)}. 108 Comparison of the zero porosity mechanical properties of YBaQCu3OM and HAP (Calo(PO4)6(OI-I)2). 112 Table 3-10 Table 3-11 Table 3-12 Table 3-13 Table 3-14 Fracture toughness/porosity trends fit to the MSA model (ln(Kw) = ln(K,c0)- kap) fiom the literature data for a variety of porous ceramics. The processing method (prc) and fiacture toughness test method (meth.) are included. Thermal expansion data for Alumina, fi‘om American Institute of Physics Handbook 2"” Edition [76]. The averaged thermal expansion coefficients for Sintered hydroxyapatite specimen numbers 4-8, and 10-12 by dilatometry. The delta is the diflereme between the cooling data and the heating data: Delta = a cool - or heat. Dielectric characteristics for Coors ADS 995 alumina substrate. Data from Tech Specs, Thin Film Substrates Technical Specifications 10-3-0897 fiom Coors Ceramics, Golden, CO [72]. Relative permittivity for Sintered hydroxyapatite fi'om this study and Fanovich et al. [78] and the fit parameters for the Maxwell-Garnet equations. 116 120 125 128 130 Figure 1-1 Figure 2-1 Figure 2-2 Figure 2-3 Figure 2-4 Figure 2-5 Figure 2-6 Figure 2-7 Figure 2-8 LIST OF FIGURES Vickers indentation impression and the characteristic dimensions used in the hardness calculation and the toughness equation (Equation 1-6), A) top view left and B) side view. A schematic of the side view of representative Palmqvist cracks is given in C) [42]. Particle size analysis from Cerac, Inc.[66] for hydoxyapatite powder lot numbers X16907T and X21949, used in this study. Mean values were 10.03 and 8.81 microns with standard deviations of 6.94 and 7.05 microns for hydroxyapatite powder lots X16907T and X21949 respectively. Pressing die for disk specimens, fabricated fiom hardened steel by the Michigan State University Physics Department Machine Shop. Rectangular die press, fabricated fi'om hardened steel by the Michigan State University Physics Department Machine Shop. Tolerances set to 0.001 ". Bolted together with 5/ 16" - 18, hex head bolts before pressing. Microwave casket assembly. Hydroxyapatite green bodies were placed on the porous alumina disk. The optical thermometer measured the temperature by fiom the light emitted through the hole 1 cm above the alumina. Microwave fumace power supply, controls and microwave cavity schematic. Apparatus for measuring the mass density of the hydroxyapatite specimens using Archimedes method. The Sintered specimens were placed into the basket for the suspended mass measurement. Electrodes and experimental apparatus for the dielectric measurements. The specimen was placed between the top and bottom electrodes. The electrodes were pressed onto the hydroxyapatite surfaces by a 5.09 cm maximum opening micrometer (not shown). Parallel circuit for Q-meter measurements. 12 29 35 36 40 42 47 53 55 Figure 3-1 Figure 3-2 Figure 3-3 Figure 3-4 Figure 3-5 Figure 3-6 Figure 3-7 Figure 3-8 Figure 3-9 Density as a firnction of sintering time and different uniaxial die pressures for hydroxyapatite Sintered in the conventional fumaces at 1300 °C. The data was taken from a subset of the data presented in Table 3-1. Density as a function of die pressure. The data was taken from Sintered hydroxyapatite specimens Sintered at 13000 C for 15-120 minutes, which are a subset of the data presented in Table 3-1. Linear fit represents the least squares fit of the data to equation 3-1. Fit parameters are slope = 0.27% relative density per 1 MPa die pressure, y-intercept = 91% relative density and the coefiicient of determination r2 = 0.997. Density as a function of sintering temperature for hydroxyapatite specimens Sintered for 60 minutes. The data was taken from a subset of the data presented in Table 3-1. Average grain size of Sintered hydroxyapatite as a function of sintering temperature and time for both conventional and microwave sintering. The data was taken flour a subset of the data presented in Table 3-1. Grain size as a function of percent relative density for Sintered hydroxyapatite. The data was taken from a subset of the data presented in Table 3-1. Micrographs of polished and thermally etched surface of Sintered hydroxyapatite specimen number 15 Sintered at 13000 C for 45 minutes, average grain size = 3.60 microns. Micrograph of polished and thermally etched surface of Sintered hydroxyapatite specimen number 45 Sintered at 1350° C for 60 minutes, average grain size = 4.59 microns. Micrographs of polished and acid etched surface of Sintered hydroxyapatite (0.1 M HCl for 1 minute), (a) specimen number 3 Sintered at 1300° C for 45 minutes, grain size =1.98 microns, (b) specimen number 8 Sintered at 1400° C degrees for 240 minutes, grain size = 7.44 microns. Polished and thermally etched surfaces of cold isostatically pressed (250 MPa) and Sintered hydroxyapatite specimen number 22 Sintered at 12000 C for 15 minutes. Average grain size = 1.99 um. 60 61 62 63 65 66 67 68 69 Figure 3-10 Figure 3-11 Figure 3-12 Figure 3-13 Figure 3-14 Figure 3-15 Figure 3-16 Figure 3-17 Figure 3-18 Figure 3-19 Figure 3-20 Figure 3-21 Polished and thermally etched surfaces of cold isostatically pressed (250 MPa) and Sintered hydroxyapatite specimen number 23 Sintered at 1250° C for 30 minutes. Average grain size = 2.06 um. Polished and thermally etched surfaces of cold isostatically pressed (250 MPa) and Sintered hydroxyapatite specimen number 20 Sintered at 1300° C for 15 minutes. Average grain size = 2.30 um. Polished and thermally etched surfaces of cold isostatically pressed (250 MPa) and Sintered hydroxyapatite specimen number 21 Sintered at 13000 C for 120 minutes, Average grain size = 2.60 um. X-ray diffraction pattern (Cu K“, radiation) for the as-received powder and Sintered hydroxyapatite (HAP) specimens numbers 6 thru 9. Additional x-ray difli'action patterns (Cu K“, radiation) for Sintered hydroxyapatite specimens numbers 11 and 12. Vickers indentation Hardness values for Sintered hydroxyapatite as a function of density. The data was taken fiom Table 3-3. Vickers Hardness plotted against the average grain size of the Sintered hydroxyapatite. The data was taken from Table 3-3. Statistical standard deviation of the Vickers hardness of Sintered hydroxyapatite plotted against the average grain size. The data was taken from Table 3-3. Vickers indentation toughness as a function of density for Sintered hydroxyapatite. The data was taken from Table 3-3. Fracture toughness as a function of average grain size of Sintered hydroxyapatite. The data was taken from Table 3-3. Standard deviation of the fracture toughness plotted against the average grain size of the Sintered hydroxyapatite. The data was taken from Table 3-3. Ratio of the Vickers indentation crack length and Vickers indentation diagonal plotted against the % relative density, for all 410 indents on the specimens with known density measured in this study. 70 71 72 74 75 79 80 82 83 84 85 86 Figure 3-22 Figure 3-23 Figure 3-24 Figure 3-25 Figure 3-26 Figure 3-27 Figure 3-28 Log relative indentation Hardness values for sintered hydroxyapatite plotted against the volume fraction porosity along with the solutions for the linear log relative hardness as a function of volume fraction porosity (Equation 1- 8, Section 1.6). Data was taken fi'om Table 3-3 for the present study and Table 3.4 for the literature data. Fracture toughness values for sintered hydroxyapatite plotted against the volume fraction porosity. Fracture toughness values were taken fiom Table 3-3 (present study) and Table 3-5 for the data from Suchanek et al. [56] and DeWith et al. [54]. Relative fi'acture toughness values for sintered hydroxyapatite as a fimction of volume fiaction porosity and the calculated linear fit. Linear fit parameters taken from Table 3-5 for the data from Table 3-3 (present study) and the fracture toughness data in Table 3-5 from Suchanek et al. [56] and DeWith et al. [54]. A section of lower volume fi'action porosity of the relative fracture toughness values for sintered hydroxyapatite as a function of volume fiaction porosity and the calculated linear fit. Linear fit parameters taken from Table 3-5 for the data from Table 3-3 (present study) and the fracture toughness data in Table 3-5 from Suchanek et al. [56] and DeWith et al. [54] . (E/H)“ as a function of the volume fiaction porosities normalize by (ED/Ho)”, the firlly dense values of the elastic modulus and hardness respectively. The decay constants bH and bE are compiled by Rice [58] for A1203, AIN, B4C, and SiO2 (Table 3-6). The bH for HAP was determined in this study, while bE for HAP was calculated in this study using the MSA model and DeWith et al.’s data [54] (Table 3-6). Calculated fi'acture toughness of the hydroxyapatite specimens in this study using the elastic modulus as function of volume fraction porosity {K,c (E(P))} (Equation 3-6, material dependent constant bE = 2.93 [54]). Crack parameter P/c”2 as a function of the volume fiaction porosity for hydroxyapatite from this study and for YBaQCu3O7,x fiom Tancret et al. [78]. The line through the YBazCu3,O7,x data is fiom the MSA fit (lnP/c3° = ln(P/cmo} bP Pp) of the crack parameter PIC” [78]. 90 96 97 105 107 109 Figure 3-29 Figure 3-30 Figure 3-31 Figure 3-32 Figure 3-33 Figm'e 3-34 Figure 3-35 Figure 3-36 Fracture toughness YBaQCu3O7,x of as a function of the volume fraction porosity determined by SENB [78] (Table 3-7), Indentation method Klc = C (Ell-I)“2 P/cm (Equation 3-3) with the data from Table 3-7, and the indentation method (Equation 3-3) with constant elastic modulus (E = 114 GPa) fiom Tancret et al. [78]. 111 The fi'acture toughness of the sintered hydroxyapatite specimens in this study calculated with C(Pp) (Equation 3-8). 115 Normalized fracture toughness parameters (fracture toughness for Ricote [48], and fracture energy for Case et al. [81]) studies in which the fiacture toughness parameters increased or decreased slightly, over a significant volume fiaction porosity range. Normalization is done by dividing the fracture toughness by the zero porosity fi'acture toughness as determined by the intercept of the linear fit. The linear fit (included in plot) calculated for the entire range of Ricote et al’s data [48]. The linear fit calculated for the Case et al. [81] data was done for specimens with < 0.30 volume fiaction porosity. The ceramics in question are calcium modified lead titanate for Ricote [48] and gadolinia for Case et al. [81]. 117 Normalized fracture toughness for sintered hydroxyapatite specimens in this study. Normalization is done by dividing the fixture toughness by the zero porosity fiacture toughness as determined by the intercept of the linear fit. The linear fit (included in plot) was calculated for hydroxyapatite specimens < 0.20 volume fiaction porosity. 118 Thermal expansion coefficient for high purity alumina determined by dilatometry, as a function of temperature and compared with the American Institute of Physics Handbook Value [76] (Table 3-6). Dilatometry data taken fiom Appendix A. 121 Thermal expansion coefficients, measured by dilatometry on heating, for sintered hydroxyapatite specimens numbers 4-8 and 10-12. Data taken from Appendix A. 123 Thermal expansion coefficient data for sintered hydroxyapatite, from cooling dilatometry. The data taken fiom Appendix A. 124 Average thermal expansion coefficient for sintered hydroxyapatite specimens numbers 4-8 and 10-12 (heating to 6000 C) as a function of the average grain size. The data was taken from Appendix A. 126 xiv Figure 3-37 Figure 3-38 Figure 3-39 Figure 3-40 Difierence in the thermal expansion coefficient for sintered hydroxyapatite, Delta = or cooling - or Ms , as a function of temperature. Dilatometry data taken fiom Appendix A. Relative permittivity data fiom sintered hydroxyapatite and the fitted Maxwell-Gamet equations MGTl (Equation 1-9) and MGT2 Equation (1-10). The data was taken from Table 3-9. Relative permittivity for the sintered hydroxyapatite specimen numbers 14, 16, 33, 41, 42 and the alumina reference material ADS995 as a function of frequency. The data was taken fiom Appendix B. Relative permittivity for sintered hydroxyapatite specimen numbers 43-46, 49 and 50 as a function of frequency. The data was taken from Appendix B. 127 131 132 133 1.0 Introduction Natural bone is a nano composite of ceramic apatites, principally hydroxyapatite and carbonate apatite, and polymeric collagen fibrils [1]. Interest in the material properties of sintered hydroxyapatite is fueled by hydroxyapatite’s applications in structural bio-implants to replace damaged bone. Natural bone, has strength and resilience that is difficult to match artificially. Hydroxyapatite is analogous to the natural ceramic in bone and is bio-active in that hydroxyapatite is non-toxic and an interfacial bond will form between implanted hydroxyapatite and living tissue [1]. Use of this bio-activity in processing mechanically functional components, with hydroxyapatite to interface with the living tissue, requires the study of the material properties of processed hydroxyapatite. The strict requirements for the material properties and quality of bio-implants are due to the high cost of replacement in terms of physical suffering and the medical care resulting fiom the invasive implant procedure. 1.1 Goals of research The primary goal of this study is to determine the effect of the microstructure of sintered hydroxyapatite on the mechanical properties of hardness, fracture toughness, and thermal expansion. The effect of the microstructure of sintered hydroxyapatite on the dielectric permittivity is also investigated. An additional goal is to catalog the rnicrostructures of sintered hydroxyapatite produced by compacting and sintering powdered hydroxyapatite. 1.2 Chemistry of sintering Sintering of hydroxyapatite allows formation of dense solid shapes, usefiil in engineering applications, to be formed from hydroxyapatite powder. Sintering requires mechanical compaction of powder, and firing until difl‘usion combines individual particles. The compaction and firing process, along with the chemistry and morphology of the starting powder, determines the microstructure of the solid material. 1.2.1 Dehydroxylation and decomposition High temperature sintering of hydroxyapatite (HAP) powders, in air will, depending on the sintering conditions, lead to the thermal decomposition of hydroxyapatite to tricalcium phosphate (TCP) and tetracalcium phosphate (Tet-CP). Thermal decomposition of hydroxyapatite is accomplished in two steps, dehydroxylation and decomposition [2, 3, 4]. The first step, dehydroxylation to oxyhydroxyapatite (OHA) (equation 1-1) is significant at temperatures greater than 800° C in air [2, 4, 5, 6]. (33109006 (0le -> C310(P04)6 (OH)2-2x Ox + tzO gas (1'1) Dehydroxylation is firlly reversible in the presence of water vapor and elevated temperatures as shown by x-ray difii‘action [4], thermogravimetric analysis [7], and by proton electrical conductivity [8]. Other experiments have shown that dehydrated hydroxyapatite, a hexagonal apatite called oxyapatite (OAP) Caw(PO4)60 E] (where E] denotes a vacancy), undergoes little or no rehydroxyalation of oxyapatite at room temperature [4,5,6]. Decomposition to tricalcium phosphate and tetracalcium phosphate occurs at temperatures > 1000 °C (equation 1-2) afier significant loss of hydroxl groups [2,3,4]. caio(PO4)6 (OH)2 ’> 2C3} (P002 + C34P209 + H20 gas (1'2) Decomposition temperature of hydroxyapatite can be lowered significantly by reduction of the water partial pressure in the sintering atmosphere [2,3,4,9, 1 0], lowering the calcium phosphate ratio [2,11,12], and by sintering with particulate additives [10,13]. At temperatures above 1120° C, dehydrated hydroxyapatite decomposes into the a-form of tricalcium phosphate and tetracalcium phosphate [14]. B-tricalcium phosphate formed at lower temperatures will also transform into (at-tricalcium phosphate at temperatures above 11200 C. The or-form of tricalcium phosphate is metastable at room temperature, the low-temperature B-forrn of tricalcium phosphate is the stable phase below 1120°c [14]. 1.2.2 Process control of phase purity during sintering Both the dehydroxylation and decomposition reactions, 1-1 and 1-2, have H20 as a product. The driving force for reactions 1-1 and 1-2 depends, in part, on the partial pressure of H20 in the furnace. The starting temperatures of the decomposition reaction as a function of water partial pressure are; 13250 C, 1426° C, 1477° C , 1485° C, 1520° C and 1565°C at water partial pressures of 0.61 kPa, 4.25 kPa, 9.81 kPa, 12.3 kPa, 33.3 kPa and 101.3 kPa, respectively [3]. The Ca/P ratio of a hydroxyapatite powder is an indicator of the high temperature stability of hydroxyapatite. A thermogravimetric study of stoichiometric hydroxyapatite (Ca/P ratio = 1.67) determined that the hexagonal structure of hydroxyapatite is stable after losing up to 75% of the hydroxyl ions in the form of water [2]. Hydroxyapatite powders, with low Ca/P ratios, show higher rates of decomposition under the same high temperature conditions than do hydroxyapatite powders with Ca/P ratios 1.67 or greater [11]. Experimental data from the literature gives a wide range of secondary phase composition of sintered hydroxyapatite, for a variety of starting powders and sintering conditions (Table 1-1). 1.3 Relevant crystal structures Crystalline phases of hydroxyapatite (HAP) and HAP decomposition products; tricalcium phosphate and tetra calcium phosphate frequently coexist in the microstructure of the sintered solid body. Interactions between phases effect the material properties of sintered hydroxyapatite solid. 1.3.1 Hydroxyapatite Hydroxyapatite is a member of a class of minerals known as apatites, the general apatite formula is Ca,(PO4)3X where X denotes (0H),, F2. or Cl [15]. Apatites have a hexagonal crystal structure with space group P63/m [15,16,17,l8]. Hydroxyapatite in the hexagonal (apatitic) structure, P63/m has room temperature lattice parameters a0 = 0.9423 nm, and co = 0.6875 nm [18]. The theoretical density of hydroxyapatite is 3.156 g/cm3 [1 5]. Under a strict calcium phosphate ratio of GNP = 1.67, high purity hydroxyapatite Table 1-1 A survey of secondary phase formation in sintered hydroxyapatite, under varied sintering conditions from the literature. Starting Temperature Time Atmosphere (1 - B - Tetra- Reference Powder (°C) (min) TCP TCP TCP Solution 1050-1350 180 Air 0 O 0 [19] grown Solution 1450 180 Air Trace 0 0 [19] grown Filter 1150-1300 180 Air 0 0 0 [20] Cake Filter 800-1000 120 HIP in Air 0 0 0 [21] Cake Comm." 1200 60 Air 0 0 0 [10] Comm." 120 60 Moisture 0 0 0 [10] ppt." 1200 600 Air 0 0 O [22] HAP 1150 1 Microwave Trace 0 0 [23] whiskers in air HAP 1200 1 Microwave 0 0 0 [23] in air ppt." 1 100 240 air 0 Trace 0 [9] ppt." 1300 240 air 0 Trace Trace [9] ppt." 1350 240 air Trace Trace Trace [9] ppt." 1100-1300 240 Vacuum Trace ----- Trace [9] Ca/P 1250 120 air 10% 75% n/a [1 1] =1.55 CW 1250 120 air 5% 35% n/a [1 1] =1 .61 Ca/P 1250 120 air 3% 0 n/a [1 1] =1.64 Ca/P 1250 120 air 0 0 0 [11] =1.67 Ca/P 1250 120 air 0 0 0 [11] =1.73 ‘- Commercial powder " — Precipitate fi'om solution. 5 precipitates into a monoclinic structure with space group P2lfb [24,25,26]. Room temperature hydroxyapatite monoclinic (P2,/b) lattice parameters are; a0 = 0.94214 nm, bo = 2a,, co = 0.68814 nm and y = 120° [24]. Monoclinic hydroxyapatite has been prepared hydrotherrnally [25], by a dry method [27], and by heating a single crystal of chloroapatite to 1200° C in steam [24]. A fully reversible phase change from the monoclinic structure to the hexagonal structure has been shown by disappearance of birefiingence at 211.5° C [25], dielectric measurements at 207° C [26], x-ray difli’action analysis and differential scanning calorimetry, with transition temperature between 2000 C and 210° C [27]. The subsequent discussion of hydroxyapatite, in this thesis, assumes the hexagonal crystalline structure, unless specifically noted otherwise. Thermal expansion coefficients along the crystallographic axes, of hydroxyapatite, reported in the literature include: a(ao) = 13.6 x1045 /° C , a(co) = 12.6 x10" /°C, varying linearly with temperature from room temperature to 850° C Trombe et al. [4]; or(a,,) = 13.1 x10“’ /°C , a(co) = 21.7 x10‘° /°C, from 22° C to 1002° C Fischer et al. [28]; and a(ao)= 12.2 x10‘° /°C , and a(co) = 11.5 x10‘6 /°C fiom 20° C to 576° C Perdok et al. [29]. The difference (Act ) in the coefficient of thermal expansion along the crystallographic axes a0 and co, from the three studies are varied: Aa = 1.0 x10"’ /°C [4], 7.4 x1045 /°C [28], 0.7 x10‘6 /°C [29]. Room temperature x-ray diffraction studies of sintered hydroxyapatite reveal the shrinking of the lattice parameter 30 as a function of sintering temperature. Lattice parameter ao decreases from 0.9416 nm, for sintering at 900° C, to 0.9411 nm, for sintering at 1450°C [19], and a0 = 0.9428 nm, for sintering at 1050° C and a, decreases to below 0.9425 nm for sintering at 1250° C [7]. The lattice parameter a0 reduction is due to the loss of (OH‘) [7,19]. The hydroxyapatite used by Trombe et al. [4] for crystallographic thermal expansion measurements was synthesized from tricalcium phosphate and calcium carbonate by heating for approximately one month at 1000° C-1100° C in a nitrogen/water vapor atmosphere. The hydroxyapatite material used in the study by Fischer et al. [28] was extensively heat treated at1200° C for 4 hours before x-ray analysis of the crystallographic thermal expansion coefficients. Perdok et al. [29] prepared stoichiometric (Ca/P ratio = 1.659 d: 0.03) hydroxyapatite crystals at 70° C. The differences in the concentration of thermal vacancies between the hydroxyapatite formed in the three studies may explain the variance in the thermal expansion along the crystallographic axes [29]. For a hexagonal crystal structure equation 1-3 is used to calculate the average linear thermal expansion coefficient (a) from the measured crystallographic therml expansion coefficients. a = [2a(a.)+ a(¢.)] / 3 (1-3) The average linear thermal expansion coefficient for polycrystalline hydroxyapatite calculated from the above lattice thermal expansion data are: 13.24 x10‘6 /°C [4], 15.97 x10“ /°C [28], and 11.95 x10‘6 /°C [29]. Jarcho et al. recorded 11 x106 /°C for the average linear thermal expansion coefficient for hydroxyapatite, at temperatures between 25° C to 225° C [30], the average linear thermal expansion coefficient for hydroxyapatite measured by dilatometry, by Choi et al. equals 16.9 x1045 /°C [31]. 1.3.2 Tricalcium phosphate The high terrrperature a-tricalcium phosphate crystallizes in a monoclinic structure with space group P2,/a, and room temperature lattice parameters a0 = 1.2887 nm, and b0 2.7280, co = 1.5219 nm, with [3 = 126.20°[32]. The theoretical density for a -tricalcium phosphate is 2.863 g/cm3[32]. The B-form of tricalcium phosphate is a rhombohedral with space group R3c, room temperature lattice parameters a0 = 1.0439 nm, and co = 3.7375 nm [33]. The theoretical density for [3 -tricalcium phosphate is 3.067 g/cm3 [33]. 1.3.3 Tetracalcium phosphate Tetracalcium phosphate has a monoclinic crystal structure with space group P2,. Room temperature lattice parameters are a0 = 0.7023 nm, and b0 = 1.1986, c() = 0.9473 nm, with [3 = 90.90°[34]. Theoretical density for tetracalcium phosphate is 3.051 g/cm3 [34]. 1.4 Microstructure of sintered hydroxyapatite The microstructure of sintered hydroxyapatite consists, in part, of directly measurable characteristics as porosity, grain size, impurity levels, and the presence of secondary phases. The porosity, grain size, impurity levels, and secondary phase relationships are controlled to engineer the processing of hydroxyapatite. Other microstructural features that can only be measured indirectly include residual stress levels in grains and microcracking between the grains. Interactions between different solid phases and pores are also integral to the description of the sintered microstructure of a solid body. 1.4.1 Residual strain energy and microcracking Thermal expansion mismatch between phases and thermal expansion anisotropy among individual grains create rnicrostresses in sintered ceramics during cooling fiom the sintering temperature [35]. Stresses induced by thermal expansion mismatch and thermal expansion anisotropy may generate internal microcracks between grains in brittle materials [3 5]. A microcrack is a separation between grains with a large aspect ratio (length to width ratio). Anisotropic crystals exhibit thermal expansion anisotropy in which the thermal expansion varies along different crystallographic axes [3 5]. Thermal expansion anisotropy of hydroxyapatite has been measured directly with varying results (section 1.3.1). The decomposition products of hydroxyapatite are all noncubic and exhibit some degree of thermal expansion anisotropy (TEA). The maximum elastic strain energy Ust stored in a single grain in a material subject to thermal expansion mismatch or TEA may be expressed as equation 1-4 [36]. Ust=1/8(E ATzAamz) (1.4) where the maximum thermal expansion difference between two adjacent grains is Aer"m , the modulus of elasticity is E, and the temperature change is AT. Equation 1-5 is an energy criterion for nricrocracking which relates grain size, grain geometries, material properties, and thermal history and is written in terms of the maximum strain energy, equation 1-4, and a critical grain size, 8e. [36]: gc,=kyf/E( ATzAamz) (1-5) where k is a constant based on the grain’s geometry and 7f is the fracture surface energy. The quantities E, AT, and Au are as defined following equation 1-4. The or to B tricalcium phosphate transformation results in a 6.7% volume decrease which contributes strain energy to the sintered solid body [32,33]. The magnitude of the strain energy for [3 tricalcium phosphate grains transformed in a hydroxyapatite matrix is a function of grain volume and the quantity and distribution of B tricalcium phosphate crystals [7,9]. 1.5 Microstructure and mechanical properties 1.5.] Indentation hardness Vickers indentation hardness is measured by the size of the permanent deformation of the surface of the tested material and the load required to produce the impression (Figure 1-1). Vickers indentation hardness can qualitatively measure the relationships between plastic deformation mechanisms, such as slip and twinning found in alumina at room temperature [37], and microstructure. 1.5.2 Indentation method to determine fracture toughness Fracture toughness is a measure of a materials resistance to crack growth [35]. Lawn and Wilshaw [3 8] reviewed the principles of indentation fracture and considered the applications to fracture toughness evaluation. Due to the complex indentation stress field considerable uncertainties exist in the absolute fiacture parameters. However, 10 determinations of the relative fiacture toughness for a material can be determined with an accuracy that is comparable to most other fracture techniques [3 8]. For indentation fracture toughness measurements sharp indenters create an intense elastic stress field about the point of indentation [3 8]. Shear and hydrostatic compression near the indenter point account for the residual impression and microfracture initiation [3 8]. Subsequent cracking is caused by tensile stress acting on the initial microfiacture [38]. Toughness is measured by the extent of the subsequent cracking[38]. Evans and Charles [39] did a pioneering study of ceramic materials with a large range of hardness, toughness and Poisson’s ratios. Evans and Charles compared fi'acture toughness values (Kc) determined by the double torsion technique with the Vickers indentation parameter (c/a), where c is the indentation crack length and a is the Vickers indent diagonal[39]. A relationship between (KC (1)) / (H I a [H/(D E]°“) and (c/a) was found for well behaved halfpenny cracks created by Vickers indentation, (Figure 1-1), where (I) is a constraint factor, H is Vickers Hardness (Pa), and E is the elastic modulus (Pa). Anstis et al. [40] further quantified the indentation toughness technique using fi'acture toughness data from traditional techniques (double torsion and double cantilever bend) compiled from a variety of ceramics with a large range of hardness, toughness and elastic modulus. Comparing toughness data measured by double torsion and double cantilever bend methods to Vickers indentation parameters leads to an equation for the fixture toughness as a function of the Vickers indentation crack length in which accuracy better than 30-40% is readily attainable (Equation 1-6) [40]. 11 A) B) I 2a 1 l 23 | \/ '<—— 2c —>' K 2c > I Vickers Indentation and fully developed half penny cracks C) l 2a I w/k/ |< >| Vickers Indentation and Pahnqvist cracks 2c Figure 1-1 Vickers indentation impression and the characteristic dimensions used in the hardness calculation and the toughness equation (Equation 1-6), A) top view left and B) side view. A schematic of the side view of representative Palmqvist cracks is given in C) [42]. 12 E H. , P K.=§( )"‘;a-e- (M) where H, is Vickers Hardness, E is the elastic modulus, P is the indentation load and indentation crack length is c. The value of the non-dimensional empirically determined constant Q is equal to 0.016 d: 0.004 [40]. Other researchers [41,42] have successfully extended the Evans and Charles/Anstis model to the low load regime where Palmqvist cracks (Figure 1-1) are encountered. 1.5.3 Hardness and fracture toughness studies Hardness and fi‘acture toughness measurements of sintered ceramics have been used to determine the influence of microstructural features such as grain size, porosity, crystallographic phase changes, grain morphology, degree of crystallinity in glass ceramics, and the elastic and plastic anisotropy. The inverse trend of hardness as a function of grain size (g.s) (Hv °< [g.s]"") for a variety of dense oxide and non-oxide materials has been studied by Rice et al. [43]. Several ceramic materials show a distinct Vickers hardness (H,) dependence on grain size[43]. A 10 fold increase in grain sizes produces; ~70% decrease in Vickers hardness (I-L) for MgO, ~30% decrease in H, for BeO, ~10% decrease in H, for A1203, and a ~5% decrease in H, for MgA1204 and B4C[43]. ZrO2 shows very little grain size dependence on H\, over the grain size range 0.5 pm to 50 um[43]. Other materials such as TiBz, SiC, TiC and Si3N4 show little effect of a grain size hardness relationship over the grain size ranges studied (10 rim-140 urn for TiB2 and SiC, 20 rim-140 am for TiC and Si3N4)[43]. 13 Krell et al. [44] investigated the hardness versus grain size relationships of high purity sintered alumina of 99.2% relative density . High purity submicron alumina powder (Taimicron-DAR) was used to sinter six specimens, with grain sizes ranging from 0.4 ,um to 4 am, to 99.2% :i:0.4% relative density [44]. Vickers hardness testing (10 N load) measured a continuous decrease in Vickers hardness as the grain size increased (H, = 22 GPa at grain size = 0.4 pm to Hv= 15 GPa at grain size = 4 um[44]). The increase in hardness as the grain size decreased for sintered alumina was attributed to the reduction of dislocation mobility as the grain size decreases [44]. Hardness and toughness measurements of sintered Y203, processed by either hot isostatic pressing (HIP) or vacuum sintering (VS), were compared using indentation hardness techniques with indentation loads ranging fiom 25-200 N [45]. Microstructural comparison showed that HIPed Y203 was denser than vacuum sintered Y203 (99.7% versus 98.8% theoretical density) with HIPed Y203 having much smaller average grain size than vacuum sintered Y203 (0.7 pm --HIP versus 15 um "vacuum sintered) [45]. HIP Y203 was ~16% harder than vacuum sintered Y2O3 (H, = 7-7.5 GPa --HIPed Y203 versus 6-6.5 GPa «vacuum sintered Y203 [45]). Direct crack measurements of the Vickers indentation cracks (indentation toughness technique) of HIP and vacuum sintered Y203 were used to calculate fi'acture toughness. The calculated fi'acture toughness of HIPed Y203 was ~9% greater than vacuum sintered Y2O3 [45]. Nano indentation of nano crystalline ZnO and nano phase TiO2 shows a decrease in hardness as the porosity of the specimen increases [46,47]. Nam-crystalline ZnO and nano-phase TiO2 exhibited an almost linear increase in hardness as a function of sintering temperature, regardless of the grain sizes[46,47]. The average grain sizes for nanophase 14 TiO2 were estimated from previous sintering studies of nanophase TiO2 from 12 nm as compacted (no sintering) to 243 nm for TiO2 sintered at 900° C [47]. The average grain sizes for ZnO is 7.4 nm (unsintered) to 140 nm for ZnO sintered at 700° C, determined by dark field TEM [46]. Relative densities of TiO2 were estimated from previous sintering studies of nanophase TiO2 and increase with increasing sintering temperatures. Relative density equals ~ 75% for the as compacted TiO2 and reaches a maximum of ~ 90% at 900° C [47]. The relative density of nanocrystalline ZnO, estimated from elastic modulus measurements, increased as the sintering temperature increased, for specimens sintered at the maximum temperature of 900° C the relative density was estimated to be 90-95% [46]. Hardness decreases as porosity increases from 1.8 GPa at 15% relative porosity to 4.0 GPa at 5 - 10% relative porosity for ZnO [46] and fi'om 1.3 GPa at 25% relative porosity to 9.0 GPa at 10% relative porosity for TiO2 [47]. Indentation hardness and indentation toughness measurements of calcium-modified lead titanate ceramics show the influence of porosity, grain size and tetragonal distortion [48]. Calcium-modified lead titanate specimens were sintered with relative porosities ranging 2% to 30%, grain sizes between 2.4 pm to 6.0 um, and three Ca/Pb ratios; 24/76, 26/74, and 35/65. Hardness decreased as porosity increased, from 3.3 GPa at 4-7% relative porosity to 2.3 GPa at 17.5% relative porosity for specimens with Ca/Pb ratio of 26/74, regardless of grain size (2.99-5.95 pm) [48]. The slopes of the hardness versus porosity curves were significantly different for the different Pb/Ca ratios signifying the effect of tetragonal distortion [48]. The quantity, morphology and distribution of fluorine containing mica embedded in a glass matrix influenced the hardness of a glass ceramic composite, due to the weakly 15 bonded cleavage planes of the mica [49]. Vickers hardness decreased as the mica crystals increased in number and the aspect ratio increases [49]. A precipitous decline in hardness is measured as the growth of mica crystals caused them to interlock [49]. Elastic and plastic anisotropy of highly textured (Bi, Pb)2 SrZCaQCu3Ox superconductor has been studied by indentation hardness measurements as a fimction of the relative orientation of the indent [50]. Maximum hardness (H, = 45-50 MPa) is measured when the Vickers indent diagonal is parallel to the grain texture and a rrrinirnum hardness (H, = 35-40 MPa) is measured when the indent diagonal is 45 degrees ofi‘ of the grain texture [50]. The fiacture toughness of dense [3 Si2N4 (98.9% :t 0.2% relative density) measured by the single-edge precracked beam (SEPB) method, is a function of the grain size and morphology [51]. Intermediate sintering times (4 hours at 2000° C) for [3 Si2N4 produces an optimal number and distribution of large grains (2-20 pm diameter and 10-300 pm in length), in a matrix of srmll grains (0.2-2 um diameter and 1-5 am long) to produce the greatest fracture toughness (ch = 10.3 MPa m'”)[51]. Specimens sintered for 2 hours at 2000° C had a microstructure with fewer large grains distributed evenly in the matrix of small grains and measured the lowest fiacturc toughness (K,C = 8.5 MPa mm)[51]. Specimens sintered for 8 hours at 2000° C exhibited an increase in the number and clustering of large grains which decreased the measured fiacture toughness to Km = 8.8 MPa m"2 possibly due to an increase in microcrack density [51]. 1.5.4 Fracture energy and microstructure In studying the apparent fiacture energy 7 as a function of flaw size Rice et al [52] 16 found that ceramics with microstresses exhibit a significant decrease in the apparent fiacture energy as the flaw sizes decrease to the scale of the intergranular microstresses (i.e. grain size) [52]. Microstresses are defined to be stresses between neighboring grains and are due to thermal expansion anisotropy (TEA), volumetric phase changes and thermal expansion mismatch between phases. As the grain size increases the critical flaw size increases and the apparent fi'acture energy decreases. Since the fiacture energy 7 is related to fiacture toughness by Kc E(E y)” [53] the fracture toughness decreases for materials with microstresses between grains as the grain size increases [52,53]. A review of fiacture energy data for several dense ceramics [53] showed a pronounced maxima in the fiacture energy versus grain size relationship with a subsequent substantial decrease in fiacture energy as the grain size increased firrther. The rnaxima in fi'acture energy at a particular grain size, were most pronounced in non-cubic materials such as A1203 at ~ 80 um, TiO2 at ~ 16 am and NbIOS at ~ 6 um [53]. The rnaxima in the fracture energy and the continued decrease in fracture energy as the grain size increased beyond the critical values has been correlated with the onset of microcracking by other measurements, including the elastic modulus as a function of grain size [53]. 1.5.5 Indentation studies of sintered hydroxyapatite Vickers hardness values, fiom Kijima et al. [19] (indentation load 0.98 N), for monophase, dehydrated hydroxyapatite (OHAP) ranged between 6.3 GPa and 6.5 GPa for specimens with grain sizes between 1.6 pm and 6.5 pm and a narrow range of density (99.0-99.4% theoretical density). A steady increase in the fiacture toughness, measured by three point bend test, as a 17 function of density is recorded for sintered hydroxyapatite made fiom a commercial powder (Merck A.G.) [54]. The fiacture toughness values ranged from K,C = 0.6 MPa in"2 at ~73% theoretical density to KC = 1.0 MPa in“2 at ~97% theoretical density [54]. As the sintering temperature increased the density, grain size and amount of HAP decomposition increased [54]. Tricalcium phosphate was detected, by x-ray diffraction, for the four densest specimens sintered at temperatures greater than 1150° C (out of six total specimens) [54]. The average grain size of the sintered hydroxyapatite ranges from 1.7 pm to 4 um [54]. The Knoop Hardness (1.96 N load) of pure and partially decomposed HAP increased as the density increased for all specimens [9]. The polyphase material (partially decomposed to a-tricalcium phosphate and tetra-calcium phosphate) exhibited significantly greater hardness as a function of density [9]. HAP sintered to 78% relative density yielded a Knoop hardness of Hk = 2.04 GPa for decomposed HAP versus Hk = 1.03 GPa for monophase HAP [9]. HAP sintered to 76% relative density exhibited a Knoop hardness of H, = 1.63 GPa for the decomposed HAP versus H. = 0.93 GPa for monophase HAP [9]. The grain size was 90 um for the HAP specimen sintered in air at 1200° C to 77.6% relative density [9]. The effect of decomposition on the hardness of the sintered hydroxyapatite specimens shown by Wang et al. [9] was explored firrther by Slosarczyk et al. [11]. The phase composition and Knoop hardness (indentation load not reported) of sintered hydroxyapatite prepared fiom powders with different Ca/P ratios (Table 1- 1) was measured [11]. Hydroxyapatite specimens (of similar density 3.06 g/cm3 to 3.11 g/cm’, and a grain size range of 6-8 pm) exhibited Knoop hardness values of Hk = 4.9 GPa (40% 18 TCP, 60% OHAP), Hk 4.6 GPa (85% TCP, 15% OHAP) and finally Hk = 4.2 GPa for hydroxyapatite specimens conrposed of 3% TCP, 97% OHAP [11]. Monophase dehydrated hydroxyapatite specimens gave a Knoop hardness of HR = 2.7 GPa (2.89 g/cm3), Hk = 4.0 GPa (3.10 g/cm’) with a maximum Knoop hardness Hk = 4.6 GPa at a density of 3. 12 g/cm3 [11]. The theoretical densities of the decomposition phases (a TCP =2.863 g/cm3 and [3 TCP = 3.067 g/cm3) are 3-9% lower than monophase hydroxyapatite (HAP = 3.156 g/cm’). The actual relative density of the polyphase sintered specimen will depend on the percentage of decomposition phases present. Fracture toughness, measured by the single edge notched bend technique (SENB), was shown to be a function of grain size for submicron hot pressed hydroxyapatite specimens [55]. HAP was hot pressed in an argon atmosphere from 1100 C to 1250° C to 98.8% to 99.5% relative density and with grain sizes from 0.2 um (hot pressed at 1100° C for 30 minutes) to 1.2 pm (hot pressed at 1250° C for 600 minutes) [55]. Fracture toughness as a firnction of grain size peaked at 0.4 pm average grain size K,C = 1.2 MPa m"z dropping to Km < 1.0 MPa in"2 at 1.2 um average grain size [55]. Three-point bend fiacture strength measurements on the hot pressed HAP, by Halouni et al. [55] displayed a more significant peak of the fi'acture strength as a finction of grain size compared to the fiacture toughness peak. The fi'acture toughness and fi'acture strength peaks were at the same grain size [55]. The fracture strength of = 137 MPa, at 0.4 pm average grain size, decreased to 0f = 99 MPa at 1.2 pm average grain size [55]. Maximums in fracture toughness and fracture strength as a function of grain size were attributed to the start of microcracking at the critical grain size, due to thermal expansion anisotropy [55]. There was no attempt to determine the presence of 19 decomposition phases in the hot pressed hydroxyapatite [55]. The fracture toughness as a firnction of processing methods and reinforcing whisker content was determined by Suchanek et al. [56], for hydroxyapatite whiskers in a hydroxyapatite matrix. Fine grained elongated hydroxyapatite powder (grain size 20-50 nm diameter 100-3 00 nm in length) were reinforced with 0-3 0% hydroxyapatite whiskers (whiskers of 1-4 mm diameter with aspect ratio of 5-35, average aspect ratio = 15). The HAP powder and HAP whiskers were mixed, pressed into pellets and sintered via three methods; pressure-less sintering, hot pressed sintering and HIP sintering. Hot pressing and HIP sintering were done in an argon atmosphere and lead to traces (< 5%) of decomposition product [S-TCP [56]. The fi'acture toughness of HAP/(HAP 0-30% whisker) composites, decreased exponentially as the porosity increased regardless of the processing method and whisker content [56]. Maximum fracture toughness values (Km = 1.4-2.0 MPa mm) were achieved by hot isostatic pressed HAP/(HAP whisker) composites with maximum densities (97% to 99.9% relative density) and whisker contents between 20% and 30% [56]. Hot pressed HAP/(HAP whisker) composites exhibited fi'acture toughness of K.C = 1.35-1.41 MPa in“2 for HAP/(HAP10-30% whisker) composites with 92.5-97% relative density [56]. HAP/(HAP 10% whisker) composites were pressure-less sintered to 71% to 90% relative density and the resultant fracture toughness values ranged fi'om ch = 0.8 MPa in“2 (71% relative density) to KIC = 1.1 MPa in”2 (90% relative density) [56]. HAP/(HAP whisker) composites, produced by Suchanek et al [56], had higher fracture toughness values, as a function of density, than unreinforced (0% whiskers) HAP. Hot pressed unreinforced HAP (0%whiskers) exhibited fiacture toughness of K,C = 1.04 20 MPa m"2 at 97% relative density, which increased to K,C = 1.35-1.41 MPa mm for HAP/(HAP 10-30% whisker) composites (92.5-97% relative density) [56]. 1.5.6 Summary of hardness and toughness studies 1.5.6.1 Hardness Hardness increases as the grain size decreases, for many sintered ceramics [43 ,44,45]. The hardness of sintered hydroxyapatite is typically unaffected over a grain size range of 1.6-6.5 pm [1 9] and obscured by porosity and the presence of decomposition phases [9,11]. Secondary phases in a sintered ceramic have been shown to significantly influence the hardness. A large decrease in hardness was measured as the amount of mica increased in a glass ceramic, a further decrease in hardness can be attributed to the morphology of mica [49]. Decomposition phases in sintered hydroxyapatite increased the hardness significantly [9,11]. Porosity is the most significant factor in determining the hardness of sintered ceramics. Hardness decreases as porosity increases in ZnO and TiO2 [46,47], calcium modified lead titanate [48] and hydroxyapatite [9,11]. 1.5.6.2 Fracture toughness There is a complex relationship between the grain size and the fracture toughness of sintered ceramics. As the grain size increases for sintered ceramics the fiacture toughness decreases slightly [48] or increases slightly [45]. For other ceramics the 21 fi'acture toughness reaches a maxirna at a critical grain size then decreases significantly at larger grain sizes [53,55]. The grain size versus toughness relationship, with a maxima at the critical grain size, is related to microstresses between grains and microcracking [53,55]. For Halouni et al. [55], however, no efion was made to detect the presence of the decomposition of hydroxyapatite when the specimens were sintered in a moisture fiee atmosphere. The fiacture toughness was compared to grain size and as the grain size increases, due to the increasing sintering temperature, the likelihood of the decomposition of hydroxyapatite, in an argon environment, increases [55]. The presence of secondary phases does not seem to influence the fracture toughness of sintered hydroxyapatite compared to the rrricrostructural features of porosity and grain size [54], or porosity and whisker reinforcement [56]. Porosity influences the fixture toughness of sintered ceramics various ways. Fracture toughness increased as porosity increased for calcium modified lead titanate ceramics [48]. Decreasing fracture toughness with increasing porosity was determined for hydroxyapatite regardless of secondary phase composition [54] and regardless of the whisker reinforcement and the presence of secondary phases [56]. 1.6 Material properties and porosity and the minimum solid area model For a wide variety of materials the porosity dependence of hardness, fracture toughness, elastic modulus, flexural strength, compressive strength and electrical and thermal conductivity have been successfully modeled by a minimum solid area model (MSA) [57,58] derived from the minimum solid area ratios. The minimum solid area ratios were calculated from (1) the idealized stacking of solid spheres, (2) idealized 22 stacking of cylindrical or spherical pores in a solid matrix, and (3) combinations of (1) and (2), as a function of volume fraction porosity[57]. The minimum solid area model relates the material property value of a porous material in relation to the fully dense property value [57] (Equation 1-7). A = A0 e 'b” (1-7) where A is the material property of the porous material, A0 is the rmterial property of the fully dense material, b is a constant derived fi'om the stacking arrangements. Equation l-7 can be linearized (Equation 1-8). ln(A)=ln(Ao)-bP (1-8) where b is the slope of the plot of the logarithm of material property (A) versus the volume fraction porosity (P). For all possible stacking arrangements of particle and pore used in the MSA model, there exists a critical volume fiaction porosity (PC) at which the bond area between pores goes to zero. Hence physical properties dr0p precipitously to zero for volume fiaction porosities greater than the critical volume fraction porosity (PC) and Equation 1-7 no longer holds [57]. Pc lies between 0.2 and 0.7 [57]. The minimum solid area model predicts continuous decreases in the physical properties of hardness, and fiacture toughness as porosity increases for materials. Fracture toughness depends less on porosity than do hardness and elastic modulus 23 [53]. In some cases an increase in fracture toughness with increasing porosity at low porosity (< 15 volume percent) occurs or there is little or no change in fiacture toughness at low and moderate porosity ranges (up to 15 volume percent porosity respectively) [53]. In contrast, strength and elastic modulus decrease continuously as in porosity increases [53]. Fracture toughness is less sensitive to changes in porosity than is hardness, which appears to be associated with either crack bridging between pores or with the interaction between porosity and the test method [53]. 1.7 Thermal expansion and microstructure A study of several ceramic oxides by Nielson and Leipold [59] found the bulk coeflicient of thermal expansion, measured up to 2200° C in air, was unaffected by the microstructural features of porosity and grain size (Table 1-2). Coble and Kingery [60] studied the physical properties of single grain size sintered alumina with variable porosity. The bulk thermal expansion coeficient, from 25° C to 1200° C, of sintered alumina was found to be independent of porosity over a wide range of porosity (Table 1-2) [60]. Table 1-2 Ceramic oxides in which the bulk thermal expansion coefficient was constant for a variety of grain sizes and porosjy. Oxide Porosity (%) Grain Size Range (um) Reference A1203 4-49% 23 um" [60] A1203 2-8% 3-66 um [59] MgOOAle3 2-9% 36-53 pm [59] MgO 2-9% 4-161 um [59] CaO 2-4% 12-111 um [59] * Single grain size for the entire porosity range. 24 1.7.1 Thermal expansion and microcracking A polycrystalline solid that contains microcracks exhibits two characteristic features in thermal expansion analysis; (1) near room temperature, the thermal expansion of microcracked materials is lower than the thermal expansion of non-microcracked material, and (2) a large hysteresis in the thermal expansion versus temperature curve between the heating data and the cooling data from a microcracked specimen [61,62,63]. Both features (1) and (2) have been experimentally documented using dilatometry for Nb205 [61], MgTi205 [62] and aluminum titanate materials [63]. Thermal expansion hysteresis between the heating and cooling curves in dilatometry studies of sintered ceramics results fiom microcrack healing at elevated temperatures as the temperature increases [64]. The microcracks re-open as the temperature decreases, due to thermal expansion anisotropy and/or thermal expansion mismatch [64]. The low values of expansion at room temperature are due to expansion of grains into the microcracks, limiting overall specimen dilation [61]. 1.8 Dielectric properties and microstructure The relative dielectric permittivity of a polycrystalline polyphase ceramic is dependent on the porosity and phase mixture of the solid [3 5]. The Maxwell-Garnet theory (MGT) is a commonly used mixing rule derived for spherical inclusions within a host matrix [65]. The MGT, included in a study by Calarne et al. [65], can be written as equation 1-9 (MGTl). Equation 1-9 gives the relative dielectric permittivity of a mixture of spherical pores in a ceramic host matrix as a function of the relative volume density of the ceramic. When air is considered to be the host matrix with additions of spherical 25 ceramic inclusions, the MGT results in equation MGT2 (Equation 1-10). k _ kce,[3 - 2R + 2ka] MGTl 1 9 "" iR+ka.(3-R)] H _ [kce,(l + 2R) + 2(1- R)] "' ‘ [kw(1- R)+ (2+ R)] MGT2 (140) where kIn and km,r are the relative dielectric permittivity of the mixture and the ceramic respectively. R is the relative volume density of the ceramic. 26 2.0 Experimental Procedure 2.1 Starting material The hydroxyapatite specimens for this study were made from 99% pure calcium hydroxyapatite (HAP) powder from Cerac Inc., Speciality Inorganics, Milwaukee, WI. Item number C-207l-1, Lot numbers X16907 (designated as HAP I) and X21949 (HAP II). The certificate of analysis, from Cerac Inc., is reproduced in Table 2-1. Table 2-1 Specific analysis (Ca percentages) and spectrographic analysis (irrrpurities) of the hydroxyapatite powder recieved fi'om CERAC Inc., Milwaukee, WI. The data is copied fiom the Certificate of Analysis attached to material. Lot number X16907T (HAP I) Lot number X21949 (HAP II) Element *% MFound “Theoretical *% “Found MTheoretical % % % % Ca 37.28 37.00 36.24 39.90 Al < 0.01 \\ <0.01 V Cr < 0.01 <0.01 Fe < 0.01 <0.01 Mg 0.08 ‘ 0.1 k Mn < 0.01 N/A Na < 0.01 ‘ N/A Si 0.08 .\ 0.1 Sr <0.01 \\ ‘ <0.01 k * Spectrographic assay ” Specific analysis or property 2.1.1 Particle size analysis Particle size analysis of the hydroxyapatite powder was performed by Cerac inc. [66] on powder from both lot number X16907T and lot number X21949. The supplied data entered in a spreadsheet and plotted as the percent change in the amount of particles passed as a firnction of particle size (Figure 2-1). 2.1.2 Ball Milling Approximately 120 grams of hydroxyapatite powder was dry ball milled in an attempt to lower the initial powder particle size before forming and sintering. Rectangular specimens R1-R5 and disk specimens numbers 13-17 were made from ball milled HAP powder. Two 250 ml nalgene bottles were each loaded with 40 g of hydroxyapatite (either HAP I or HAP II) and 130 g of almnina grinding media (HDAL 025). Another 250 ml nalgene bottle was loaded with 220 g of zirconia grinding media (Zircon 025C) with 60 g of HAP II. The nalgene bottles were tightly sealed with nalgene screw tops. The nalgene screw tops were secured to the bottle with duct tape. All three nalgene bottles, loaded with HAP and grinding media, were tumbled at 20-40 rpm in a large rock tumbler (Thumlers, Tumbler, Model B, Tru-Square Metal Products, Auburn, WA) for 293 hours. The HAP powder in the bottles adhered to the side and formed a solid mass during tumbling. Squeezing the sides of the bottle and shaking for several seconds freed up most of the powder fiom the walls. The milling media was filtered fiom the powder using a small kitchen sieve (~2 mm sieve). The ball milled powders are denoted as HAP Ia, HAP IIa and HAPIz. The “a” sufix denotes alumina ball milling and the “z” suffix denotes zirconia ball milling. All other specimens were made from as received hydroxyapatite powder. 2.2 Specimen preparation Thin, dense hydroxyapatite disks (specimen numbers 1-50) and rectangular bars 28 +. 4—1- as» g X16907T .l: .. O _ 392 .l- X21949 1-IH— 0 :e:::: ‘ :::::::: .. 0.1 1 10 100 Particle Size (Microns) Figure 2-1 Particle size analysis fi'om Cerac, Inc.[66] for hydoxyapatite powder lot numbers X16907T and X21949, used in this study. Mean values were 10.03 and 8.81 microns with standard deviations of 6.94 and 7.05 microns for hydroxyapatite powder lots X16907T and X21949 respectively. 29 (specimens R1 -R10) were produced fi‘om powder, by die pressing and firing (Table 2-2, Table 2-3, Table 2—4, and Table 2-5). Specimens were polished and sectioned to prepare them for further measurements. 2.2.1 Die forming Hydroxyapatite powder, with no added binder, was uniaxially pressed (Carver Laboratory press, Model C, Fred S. Carver Inc., Menominee Falls, WI) in a cylindrical or rectangular hardened steel die. Applied pressures varied fiom 6.55 MPa to 24.0 MPa. The disk and rectangular dies were constructed by the Michigan State University Physics Department machine shop (Figures 2-2 and 2-3 respectively). The cylindrical steel die produced disks 5.095 cm in diameter. The rectangular die formed bar specimens 8.89 cm long and 1.27 cm wide. Afteruniaxial pressing, the steel die was cleaned with acetone and lint fiee paper towels (Kimwipes EX-L, Kimberly Clark, Roswell, GA). No lubrication was used on the steel dies. 2.2.2 Cold isostatic pressing Disk specimens numbers 20-23, 31-35 and r8-r11 rectangular bars were cold isostatically pressed (CIP) in a room temperature isostatic press (180 Spectrum Inc., Columbus, Ohio) after initial uniaxial die pressing (6.55 MPa disk specimens and 7.88 MPa for rectangular specimens). The die pressed hydroxyapatite green bodies were placed in latex condoms to keep them from being contaminated by the water medium in the isostatic press. The air in the loaded condoms was removed using a 30 cc syringe and a blunt needle to withdraw the air. Cotton thread, tied around the opening of the condom 30 Table 2-2 Processing parameters for Hydroxyapatite disk specimens. Fired in automatically controlled tube fumace. HAP I is lot number X16907 from CERAC inc. Milwaukee, WI. Specimen Material Start mass Pressure Sintering Sintering label (g) (MPa) temperature (° C) time (min.) 1 HAP I 8 6.55 1300 60 2 HAP I 8 6.55 1300 75 3 HAP I 8 6.55 1300 45 4 HAP I 8 6.55 1300 90 5 HAP I 12 6.55 1300 105 6 HAP I 12 6.55 1300 240 7 HAP I 12 6.55 1400 120 8 HAP I 12 6.55 1400 240 9 HAP I 12 6.55 1300 300 10 HAP I 12 6.55 1300 420 11 HAP I 12 6.55 1300 540 12 HAP I 12 6.55 1300 660 31 Table 2-3 Processing parameters for rectangular HAP Specimens, fired in C+M box furnace under manual control. HAP II is from lot number X21949 from CERAC. HAP Ia, HAP IIa- hydroxyapatite powder ball milled with alumina pellets, HAP Iz- hydroxyapatite powder ball milled with zirconia pellets. Specimen Material Start mass Pressure Sintering Sintering label (g) (MPa) temperature (°C) time (min.) R1 HAP la 6 23.64 1300 60 R2 HAP 12 6 23.64 1300 45 R3 HAP 113 6 23.64 1300 45 R4 HAP 11a 6 19.70 1300 15 R5 HAP Ila 6 19.70 1250 15 R6 HAP II 6 19.70 1300 15 R7 HAP I 5 19.70 1250 15 R8 HAP II 5 250 (CIP)* 1300 15 R9 HAP II 5 250 (CIP)* 1300 120 R10 HAP 11 '5 250 (CIP)* 1200 15 * Cold Isostatic Press, 150 Spectrum, Columbus OH. 32 Table 2-4 Processing parameters for disk HAP Specimens, fired in C+M box furnace under manual control. Specimen Material Start mass Pressure Sintering Sintering label (g) (MPa) temperature (° C) time (min.) 13 HAP 11a 10 24.00 1300 45 14 HAP 12 10 24.00 1300 60 15 HAP 12 10 24.00 1300 45 16 HAP Ila 10 17.46 1300 15 17 HAP Ia 10 17.46 1250 15 18 HAPH 10 17.46 1300 15 19 HAPI 10 17.46 1250 15 20 HAP H 9 250 (CIP)* 1300 15 21 HAP II 9 250 (CIP)* 1300 120 22 HAP H 9 250 (CIP)* 1200 15 23 HAP H ‘9 250 (CIP)"' 1250 30 24 HAP II 9 6.55 1200 15 31 HAP I 6 80 (CIP)" 1200 60 32 HAP I 6 80 (CIP)" 1350 60 33 HAP II 6 80 (CIP)" 1100 60 34 HAP H 6 80 (CIP)" 1250 601 35 HAP II 6 80 (CIP)" 1150/1300i 60/601' 36 HAP I 6 13.09 1100 60 37 HAP I 6 19.64 1100 60 38 HAP II 6 6.55 1100 60 39 HAP II 6 26.18 1200 60 40 HAP II 6 13.09 1200 60 it -+++ Cold Isostatic Press, [50 Spectrum, Columbus OH. Cold Isostatic Press run to 80 MPa three times. Furnace run interrupted at ~ 970 ° C, restarted at ~ 440 ° C. Specimens were cupped after first firmace run at 1150 ° C. Specimens re-fired at 1300 ° C for 60 mor minutes. 33 Table 2-4 Continued Specimen Material Start mass Pressure Sintering Sintering label (g) (MPa) temperature (° C) time (min.) 41 HAP II 6 6.55 1200 60 42 HAP H 6 13.09 1250 601 43 HAP H 6 19.64 1250 601 44 HAP II 6 26.18 1250 601 45 HAP II 6 6.55 1350 60 46 HAP H 6 13.09 1350 60 47 HAP 11 7 6 26.18 1350 60 48 HAP II 6 13.09 1 150/13001' 60/60'1‘ 49 HAP II 6 19.64 1150/1300? 60/601‘ 50 HAP II 6 26.18 1 150/1300'1' 60/601‘ I Furnace run interrupted at ~ 970 ° C, restarted at ~ 440 ° C. ‘1 Specimens were cupped after first furnace run at 1150 ° C. Specimens re-fired at 1300 ° C for 60 mor minutes. Table 2-5 Processinflarameters for disk HAP S aecimens fired in microwave cavity Specimen Material Start mass Pressure Sintering Sintering label (g) (MPa) temperature (°C ) time (min) 25 HAP H 9 6.545 1050 30 26 HAP H 9 6.545 1100 30 27 HAP II 9 6.545 1150 30 28 HAP I 8 6.545 1200 30 34 1— 12.7 cm —1 V - — sum... é ml... W | i 4 — 5.095 cm Z Plungers 5.095 cm diameter Top View 1—12.7 cm ——I Figure 2-2 Pressing die for disk specimens, fabricated from hardened steel by the Michigan State University Physics Department Machine Shop. 35 I — 8.89cm— I - V rut" a, 10.16 cm Exploded View Side Plates 16.51 cmx 10.16 cm x 3.81cm E Plungers 8.89 cm x 6.35 cm x 1.27 cm ............... Spacer Plates 10.16 cmx 3.81 x 1.27 cm . Bolt Holes for 5/16"-18 hex head bolts Figure 2-3 Rectangular die press, fabricated fi'om hardened steel by the Michigan State University Physics Department Machine Shop. Tolerances set to 0.001". Bolted together with 5/ 16" - 18, hex head bolts before pressing. 36 and the needle, was pulled tight as the plunger, on the syringe, was pulled back and the needle was removed from the condom. A second condom was pulled over the material in the first condom and again the air was removed by the syringe and needle. The outside condom was sealed with cotton thread. The double sealed specimens were then secured in the middle of a steel mesh (~3-5 mm mesh) cylinder (10 cm diameter and ~ 30 cm long), which was closed at one end, to keep them fiom being sucked in to the relief valve when the pressure was released. The mesh cylinder, with the HAP specimens tied with cotton thread to the inside, was placed in the bore of the cold isostatic press, and the specimens were isostatically pressed at room temperature. The specimens pressed at 250 MPa were brought up at 50 MPa/minute and the pressure was slowly released after a minute 250 MPa. The specimens pressed at 80 MPa were brought up and down fiom atmospheric pressure to 80 MPa three times due to equipment failure (the CIP was set to go to 250 MPa but failed). Afier removing the specimens fiom the isostatic press the condoms were cut with scissors and peeled off the ceramic bodies. The cold isostatically pressed green bodies were cleaned with acetone and Kimwipes, to remove latex and lubricant fiom the HAP specimens, before firing. 2.2.3 Conventional furnace sintering Specimens numbersl-12 were fired in a conventional electric tube firmace (Thermtec, Lenton Thermal Designs). The green bodies were placed on a 30 cm long alumina semicircular hollow platen which fit into the 8 cm diameter horizontally aligned alumina furnace tube. A light, dusting (100-200 mg) of alumina powder (AKP-SO, Sumitoma, Chemical America Inc, New York, NY) was spread beneath the green bodies 37 on the alumina platen. The alumina powder was used to prevent the bonding of the hydroxyapatite specimen to the alumina platen at high temperatures. The alumina platen and HAP specimens were centered horizontally in the firmace tube near the furnace control thermocouple. The maximum temperature, dwell and ramp rate were input into the process control unit (Model BS4937, type R Eurotherm). Heating and cooling rates were programmed for 10°C/minute to and from room temperature. Additional HAP specimens (numbersl3-20, and numbers 26-50) were fired in a box furnace C + M Inc., High Temperature Furnaces, Bloomfield, NJ). Pressed hydroxyapatite specimens were set upon a 10 cm diameter alumina platens (3 mm thick) and placed in the center of the floor in the box fumace. A light dusting (100-200 mg) of alumina powder (AKP-SO, Sumitoma) was spread between the HAP specimens and the platen. The box furnace temperature was controlled manually by adjusting the element current and using a type R (platinum/platinum-l 3% rhodium, Omega Engineering, Stamford CT) thermocouple for feedback. The thermocouple bead was placed in the center of the interior of the box furnace at floor level. Ceramic bead insulators, 80% mullite 20% glass from, Omega Engineering (Omegatite 200), insulated the thermocouple wire to the cold junction compensator box (Type R, Omega Engineering). The thermocouple leads and insulation exited the firmace interior through grooves ground into the refiactory at the fumace door. The heating rate averaged 10-20°C/minute fiom room temperature. The cooling rates were manually controlled to less than 20 °C/mir1ute down to 500 °C. Below 500 °C, the furnace interior was convectively cooled to room temperature. A Fluke 77 multimeter was used to read the voltage from the cold junction corrrpensator box output, on the millivolt scale. Millivolt to temperature conversion was 38 achieved using the “Revised Thermocouple Reference Tables” for type R (N .I.S.T. Monograph 175 Revised to ITS-90), published by Omega Engineering [67]. 2.2.4 Microwave sintering Microwave sintered specimens (numbers 25-28) were placed on 5 mm thick porous alurrrino-silicate refiactory disk 7.4 cm in diameter (Zircar Fibrous Ceramics, Zircar Products, Florida, NY). The specimen and almnina disk were then placed on the bottom of a cylindrical microwave sintering casket, Figure 2-4. The ends of the cylindrical casket were made up of 2 cm thick top and bottom alumino-silicate insulating board , SALI type (Zircar Fibrous Ceramics), cut into disks 10.5-11 cm in diameter. The body of the refiactory casket was made fiom 1.25 cm thick x 3 cm tall cylinder of low density yttria stabilized zirconia (ZYC', Zircar Fibrous Ceramics). A 5 mm diameter circular view hole was bored through the zirconia cylinder body centered 8 mm fi'om the bottom edge of the low density stabilized zirconia cylinder. The cylindrical casket was centered in the microwave cavity (model CMP-250, Wavemate Inc., Plymouth, MI), with the view hole facing the center of the microwave cavity viewport. The microwave cavity was powered by a 2kW, 2.45GHz power supply (Sairem, microwave power supply, Vaux-en-Velin, France), in the resonant mode. Microwave power was controlled manually, with a ranrp rate of 20—100 watts every 3 minutes. The temperature of the specimen was unknown below 500 ° C. The temperature of the specimen was measured by an optical thermometer (Accufiber Model 10 Optical Fiber Temperature Control System, Accufiber Division, Luxtron Corporation, Beaverton, OR) when the temperature was greater than 500 ° C. 39 Sali refiactory board 2 cm Low Density Zirconia cylinder I 3 cm 0.5 cm diameter T . 0 . I 1 gm] View hole for Optical Thermometer Porous L____J+-3.2-0.3 cm thigk specimen Alumina iESi!!EEEiii:§§§§§§§§§§§§§§!!!! 0,5 cm Sali refractory board 2 cm Exploded Side View Sali refractory 11cm Top View Figure 2-4 Microwave casket assembly. Hydroxyapatite green bodies were placed on the porous alumina disk. The optical thermometer measured the temperature by from the light emitted through the hole 1 cm above the alumina 4o Microwave forward power was adjusted and tuned to bring the specimen to the desired temperature (between 1050-1200 ° C). The microwave cavity was manually tuned by minimizing the reflected power, accomplished by adjusting the locations of the cavity short and launch probe in the microwave cavity (Figure 2-5). Tuning the cavity short and the launch probe was done when the reflected power increased more than 20% of the previous value. The reflected power was measured at the directional coupler in the microwave transmission line (Microlab/FXR) by a power meter (Model 432A, Hewlett- Packard, Palo Alto, CA). Stepper motors controlled by an IBM XT computer moved the launch probe and the cavity short. Temperature measurements were made using the optical pyrometer measuring the light radiating from the heated specimen through the view hole bored in the side of the microwave casket and passing through the microwave cavity viewport. After sintering, cool down was achieved manually at 10°C/minute to 500°C by manipulating the microwave power. Below 500°C cooling was convective. Tables 2-2, 2- 3, 2-4, 2-5 and 2-6 list the forming and sintering data for all the specimens included in this study. 2.2.5 Sectioning the sintered specimens Sintered HAP and alumina specimens were sectioned with a slow speed cutting saw (Leco Corporation, St. Joseph, MI) using a 10.16 cm diameter 0.3 mm thick diamond saw (Superabrasive grinding wheel-M4D220-N75M, Norton Company, Worcester, MA), with an oil based lubricant (VC-50, Leco Corporation). The HAP and alumina specimens were then cleaned, by hand, with a liquid detergent (Liquinox, Liquinox Co., Orange, CA) diluted ~10:1 with tap water, then rinsed with copious amounts of tap water. All HAP 41 Transmission Line Microwave Cavity Microwave Viewport Probe Directional Coupler — Water Cooled Reflected Power Meter 0.1 2450 Mhz 2000W = IBM XT Computer Power Supply Figure 2-5 Microwave furnace power supply, controls and microwave cavity schematic. 42 sintered sections were further cleaned with Kimwipes and acetone. 2.2.6 Specimen polishing Hydroxyapatite specimens and parts of specimens were polished in batches that included 3-10 specimens or parts of specimens of similar thickness (within 20%). The specimens were glued to a flat 1.91 cm thick 19.2 cm diameter aluminum disk with thermoplastic (thermoplastic cement, number 40-8100, Buehler LTD, Lake Bluff IL). The aluminum disk was placed on a hot plate (Cole-Parmer Model number 465 8) and heated to melt a 1-3 mm thick layer of thermoplastic. Next the specimens were placed on the liquid plastic and the assembly removed from the heat. An identical aluminum disk was placed on top of the specimens, to press the specimens against the lower aluminum plate until the thermoplastic set. The polishing machine (Vari-Pol VP-50, LECO Corporation, St. Joseph, MI), used two rotations and a swinging motion (random orbit) to polish the specimens. Size graded diamond grit (Warren Diamond powder company, Olyphant, PA) was applied to grit size specific polishing cloth (LECO Corporation) affixed to a rigid metal disk. The cloth covered disk and the aluminum disk, with which the specimens were glued, were loaded into the polishing machine. The polishing cloth was lubricated with 10-20ml of Ethylene Glycol solution (Microid diamond compound extender, LECO Corporation). The pressure applied on the hydroxyapatite specimens, set by an allen wrench, was set by 1/4 turn tightening of the wrench The cloth covered metal disk rotation rate was set to 150 rpm. Seven or eight diamond grit sizes on separate cloth covered metal disks were used, sequentially from the 43 largest grit to the smallest grit, to polish the specimens. Polishing paste containing diamond grit was graded for 67, 35, 25, 17, 15, 10, 6, and 1 micron sizes. Initial polishing at grit sizes of 67 or 35 micron polishing evens the surfaces, and for hydroxyapatite specimens required 40-120 minutes. For the intermediate grit sizes (25, 17, 15, and 10 micron) polishing took 15-30 minutes for each size grit. Final polishing of the hydroxyapatite specimen surface was done for 20-30 minutes with 6 micron diamond grit and 30-40 minutes with 1 micron diamond grit. Between grit size changes and after final polishing the aluminum disk with the aflixed specimens was removed fiom the polishing machine, and the specimen surfaces were cleaned with liquinox soap and a soft bristle brush. The disk and the specimens were then rinsed thoroughly with tap water and dried with paper towels. 2.2.7 Surface etching Thermal etching, in order to measure the average grain size, was performed on hydroxyapatite specimens (numbers 6 - 10 and 12 - 23). Small sections (approximately 3mm x 3mm) of polished hydroxyapatite were arranged on the alumina platen with a light dusting of alumina powder (AKP-SO, Sumitoma), to keep the specimens fi'om bonding to the alumina plate at high temperature. The specimens were placed in the box fumace and brought to 1300° C for 1 hour. Heating and cooling rates were approximately 10 °C/minute to and fi'om 13000 C. Sketches of the specimen shapes and arrangements on the plate were made for specimen identification upon retrieval from the furnace. Acid etching, for grain size measurements, was done on hydroxyapatite specimens (numbers 1-5, and 11), using 0.1 M hydrochloric acid at room temperature. Small polished sections (approximately 3mm x 3mm) were set in a 50 ml glass beaker containing 15-20 ml of 0.1 M HCl, with nalgene tweezers, for 1 minute. The specimens were removed from the acid with nalgene tweezers, placed in a 100 ml nalgene beaker full of tap water and rinsed with a large volume of tap water. 2.2.8 Specimen mounting for scanning electron microscopy Prepared hydroxyapatite specimens (sectioned and polished and either etched or indented) were attached to an aluminum disk (SEM specimen mount) with carbon tape (Structure Probe Inc., West Chester, PA) with the specimen surface of interest facing upwards. The surface of interest of the specimens was coated with ~ 21 nm of gold by a plasma coating machine (Emscope, SC500). Coating parameters set for the Emscope plasma coating machine were 3 minutes coating time (plasma on), with a current of 20 mA. The deposition rate was 7 nm/minute, at 20 mA, according to the manufacturer’s specifications [68]. 2.3 Microstucture measurements The sintered specimens were initially characterized for density, grain size, and crystalline structure, which are grouped as process variables. Further measurements of hardness, toughness, thermal expansion, and dielectric properties were made on the specimens and related to the process variables. 2.3.1 Mass density Density measurements were made by Archimedes method on whole or sectioned 45 sintered hydroxyapatite specimens. The mass of the specimen dry, wet (by distilled water), and suspended in distilled water were measured three times, and these measurements are averaged. Dry specimens were measured, to 0.1 mg precision, on a electronic laboratory scale (Accutron, Denver Instrument Co. Arvada, CO). Suspended measurements were made by measuring the mass of the specimen minus the weight of the distilled water displaced by the specimen (Figure 2-6). The wet specimens were measured after the suspended measurement, the surface water was wiped off with damp Kimwipes before measuring. The difference in the mass of the wet specimen and the suspended specimen is the mass of the displaced distilled water and can be converted into a volume using 1 g = 1 cm’. The dry mass was then divided by the volume of the displaced water to determine the specimen density (Equation 2-1). density = (dry mass) / [(wet mass - suspended mass)/(l g/cm3)] (2-1) The specimen density was then divided by the theoretical density of the hydroxyapatite (theoretical density of hydroxyapatite = 3.156 g/cm’), then multiplied by 100 to obtain the relative density as a percent of theoretical density. 2.3.2 Grain size measurements Grain sizes were calculated from scanning electron microscope (Hitachi S-2500C) micrographs of polished, etched and gold plated surfaces of sintered hydroxyapatite specimens. One or more images were taken at magnifications of 2000X or greater with Polaroid 55 fihn (Polaroid Corporation, Cambridge, MA). Twenty to thirty more lines of 46 SAMPLE Figure 2-6 BALANCE PAN \ DISTH.LED WATER Apparatus for measuring the mass density of the hydroxyapatite specimens using Archimedes method. The sintered specimens were placed into the basket for the suspended mass measurement. 47 5 cm were superimposed on the rrricrographs and the grain boundary intercepts along these lines were counted. The total length of the superimposed lines was divided by the number of intercepts multiplied by the magnification in order to determine the average grain size of the specimen (Equation 2-2). L, . = ' 0X 2-2 gs NIOM ( ) where g.s. = Grain size (meters), Ln = Total length of the drawn test line(s) (meters), NI = Total number of grain boundary intercepts, M = Magnification of image, X = Geometric constant (1.0 was used for HAP). An excess of 200 intercepts/image were used in the calculations. 2.3.3 X-ray diffraction Unsintered hydroxyapatite powder (HAP I fiom Cerac Inc.) and sintered hydroxyapatite powder were pressed into a ~ 3/16 “ diameter ground out depression in a glass microscope slide. The glass microscope slide was ground out with a coarse SiC grit impregnated 1/ " diameter brass ball on the end of a 5 - 3000 rpm variable speed rotary tool (Sears and Roebuck and Co., Hoflinan Estates, IL) . The sintered powder was obtained fiom the sintered specimens numbers 6-9, 11 and 12 by crushing and grinding a small (approximately 4 mm x 4 mm) section using a Coors porcelain mortar and pestle (mortar number 60316, pestle number 522-4, Coors Ceramics Company, Golden, CO). The grinding was done until the powder was very fine and adhered to the side of the mortar. About 10-20 turns of the pestle were required to produce appropriate powders 48 once the specimen was broken up. The glass slide with the hydroxyapatite powder sample was installed into the target holder in the SCINTAG 2000 x-ray machine (Model number XDS 008, serial number 091, Scintag USA inc.). The powder sample was then irradiated with a parallel beam of CuKal. radiation (A = 15.405nm) collimated from the x-ray source by 2mm and 2mm slits ( 40 mm apart) and then a nosepiece 54 mm long with a 1.5 mm diameter hole. The x-rays were collimated after the sample by 0.5m and 0.3rmn slits 40 mm apart before the x-ray detector. A step scan program was run at 0.2 min/degree with a step of 0.02 degrees. The counts versus dim‘action data were stored in the controlling computer (various PC’s, with proprietary software fiom Scintag). A Scintag program converted the binary data into ASCII which was then downloaded into a spreadsheet (Quattro Pro 8, Corel Corporation, Ottawa, Ontario, Canada) for analysis and plotting. The plane spacings for each specimen were calculated using the wavelength fiom CchrL, the 20 peak from the three major difl'raction peaks on the difli‘action curve, and the Bragg law (Equation 2-3). A = 2 d sinO (2-3) From the plane spacing the lattice parameters (a and c) were calculated using the plane index for the three major difli'action peaks, from the ASTM powder diffraction files for inorganic phases [69], and the {plane spacing equation for the hexagonal lattice fiom appendix 3 in Cullity [70] (Equation 2-4). d2‘3 1 4 h2 + hk + k2 I2 02 + 67 (2'4) 49 2.4 Hardness and fracture toughness measurements Hardness and toughness measurements were made by measuring Vickers indentations on the polished faces of the hydroxyapatite specimens (V ickers micro hardness tester, Buehler LTD, Lake Blufi‘ IL). Loads of 2.94 N, 4.90 N, and 9.81 N were used with a 10 second loading time and a head speed of 70 uni/s. To guarantee 7 tolO recordable indents, 10 to 20 indentations per specimen were made. Indentations with surface spalling were rejected. For the hardness calculation the diagonal lengths of the indent (2a), in meters, were recorded (Figure 2-7). The total length of the radial cracks plus the diagonal lengths of the indent, (2c), were measured and recorded for the toughness calculation. The hardness and toughness of the sintered HAP specimens were calculated using Equations (2-5) and (2-6) and‘the Vickers indentation measurements. These equations and the empirical constant C were from Antsis et al. [40] presented by Wachtrnan [71]. The hardness H is calculated in GPa by Equation 2-5. H = P / 232 (2-5) where P is the load in Newtons, a is 1/2 the length of the indent diagonal in meters. Fracture toughness K,C was calculated using Equation 2-6. Kic = C (E/HY'ZP 0-3/2 (2-6) where C = 0.016, E is the elastic modulus (115 Gpa [12]), H is the hardness calculated in 50 (2-4), P is the load in Newtons and c is 1/2 the indent crack length in meters. 2.5 Thermal expansion 2.5.1 Specimen preparation for thermal expansion measurements Thermal expansion measurements used hydroxyapatite specimens numbers 4-8, 10-12 and a sintered alumina reference specimen (starting powder AKP 30-4 Sumitoma) microwave sintered at 15000 C for 30 minutes by KiYong Lee, a colleague in the Material Science Department at Michigan State University. The specimens were cut into rectangles approximately 10mm x 5mm x 3m with the slow speed diamond saw. The long axis opposing faces were ground parallel by hand sanding with wet 320, 400 and 600 grit wet/dry silicon carbide paper (3M corp. Minneapolis MN). Final surface polishing of the opposing faces was done by hand on the Vari-Pol polishing machine at 150 rpm with grit sizes starting at 17 um, then 15 um and finally 10 pm, for 2-3 minutes at each grit size. Hydroxyapatite specimens numbers 4-8, 10-12, and the alumina specimen were then annealed at 9000 C in a box furnace (Type 59344, Lindbergh Blue, Asheville, NC) for 13 hours. 2.5.2 Thermal expansion measurement Hydroxyapatite and alumina specimens were placed one at a time in the specimen well of the thermomechanical analyzer (DuPont Instruments, Model 943) so that the measurement was to be made on the long axis. The thermomechanical analyzer was programed to heat the specimen at 10°C/minute to 800° C, hold at 8000 C for 0, 10, 20, 51 30, or 60 minutes. Thermal expansion data were recorded continuously during heating to 8000 C and the cooling to room temperature. Thermal expansion coefficients were calculated, in uni/m ° C, at 25° C intervals and the data was inserted into a spreadsheet. 2.6 Dielectric measurements 2.6.1 Specimen preparation for dielectric measurements Disk specimens with both surfaces parallel and polished to 1 micron (see section 2.2.6) and at least a 2.22 cm diameter continuous surface were used for the dielectric measurements. 2.6.2 Dielectric apparatus and measurements Quality factor (Q), and parallel capacitance (Cp), measurements of sintered hydroxyapatite specimens (numbers 14, 16, 33, 36, 41-46, 49, 50) and an alumina substrate material (ADS 995, Coors Ceramics Company, Golden CO) were made with a Q-Meter (Model 4342A,Hewlett-Packard, Palo Alto, CA). The alumina substrate material (ADS 995) was used as a reference material since the dielectric properties are well documented [72]. A guarded electrode system (similar to the ASTM standard- D 150-95 [73]) was constructed fiom rigid copper and copper pipe (Figure 2-7). The copper electrodes, copper guard ring and the G10 (Glass fabric and epoxy resin system) were cut on a lathe. The contact surfaces of the copper electrodes and guard ring were polished with 25 um diamond paste on the Leco random orbit polishing machine, using a specimen holder designed to polish parallel surfaces. The rigid copper electrodes were 52 Top G-10 [ 2.22 cm Cu l 2.22 cm |__2.22cm _| —“l 1.905 cm Cu “I Bottom G-10 \ Side View 0.16 cm air gap Electrode surface for both Electrodes I FIND Q- Meter l Figure 2-7 Electrodes and experimental apperatus for the dielectric measurements. The specimen was placed between the top and bottom electrodes. The electrodes were pressed onto the hydroxyapatite surfaces by a 5.09 cm maximum opening micrometer (not shown). 53 held in place and compressed on the ceramic specimen by a 2 inch micrometer (China). The dielectric measurements were made using the “Parallel Connection” for high impedance components using reference inductors for each frequency range (reference inductors 16471A through 16490A, Hewlett-Packard). The parallel connection circuit was used (Figure 2-8). Resonance was first established in the circuit without the unknown Z. The Q (Q,) and the capacitance (C,) are recorded. The unknown Z (specimen in the electrode apparatus) is then connected and a new resonance established in the circuit. The new Q (Q) and capacitance (C2) were recorded. The effective Q, and the effective parallel capacitance Cp were then calculated (Equations 2-7 and 2-8). Q = Q1Q2 (C1 ‘ C2) /(AQ C1) (2'7) C. = (C. - C.) (2-8) Three capacitance measurements and three Q measurements were taken at 45 kHz for the hydroxyapatite specimens and the ADS995 alumina substrate. The averages of the three capacitance measurements and three Q measurements taken at 45 kHz were then used to correct for the capacitance of the circuit using the calculated capacitance of the ADS995 alumina substrate. Additional measurements were taken at 8 higher frequencies, from 72.2 kHz to 7.3 MHz (one measurement per fiequency) at room temperature for the hydroxyapatite specimens and the Coors alumina substrate ADS 995. 54 Hi Hi STABLE COIL nknown Q-meter Q Low Figure 2-8 Parallel circuit for Q-meter measurements. 55 3 Results and Discussion 3.1 Sintered hydroxyapatite microstructure Processing parameters for sintered hydroxyapatite were varied to allow for a wide range of density and grain size of the sintered microstructure (Table 3-1). 3.1.1 Sintered density For the uniaxially die pressed conventionally sintered hydroxyapatite specimens, die pressure has the most significant influence on the final density (Figure 3-1). A linear relation between die pressure and final sintered density was calculated. The relative density for sintered hydroxyapatite specimens sintered at 13000 C for less than 120 minutes (numbers 1-5, 13-16 and 18, see Table 3-1) were linear when plotted as a function of the applied uniaxial die pressure (Figure 3-2). Linear regression analysis (Quattro Pro, Corel , Ontario, Canada) determined a SIOpe of 0.27% relative density for every 1 MPa of applied pressure. The coefficient of determination r2 = 0.997 for the regression analysis. The densification rate for hydroxyapatite sintered for 60 minutes was significantly enhanced at temperatures greater than 11000 C (Figure 3-3). 3.1.2 Sintered grain size Average grain size as a firnction of sintering temperature and sintering time shows significant scattering (Figure 3-4). This is partially due to the differences in the ramping rates and furnace geometries for the two conventional furnaces used in this study. Ramp rates were electronically controlled in the tube furnace to 10° C/min and manually controlled in the box furnace to 5-20° C/min. The grain size-sintering temperature Table 56 3-1 Processing parameters, density and grain size data for sintered hydro xyapatite Specimen Pressure Sintering Sintering Furnace Density Grain size label (MPa) time (min) temperature type (g/cm’) (microns) (°C) 1 6.55 60 1300 Tube" 2.88 1.68 2 6.55 75 1300 Tube“ 2.88 1.65 3 6.55 45 1300 Tube“ 2.88 1.98 4 6.55 90 1300 Tube" 2.89 2.07 5 6.55 105 1300 Tube“ 2.88 2.14 6 6.55 240 1300 Tube" 2.92 2.32 7 6.55 120 1400 Tube“ 3.05 5.65 8 6.55 240 1400 Tube“ 3.06 7.44 9 6.55 300 1300 Tube" 2.97 5.93 10 6.55 420 1300 Tube" 2.95 5.93 l 1 6.55 540 1300 Tube“ 2.98 6.73 12 6.55 660 1300 Tube“ 2.79 6.80 13 24.00 45 1300 Box" 3.04 4.37 14 24.00 60 1300 Box" 3.00 3.70 15 24.00 45 1300 Box" 3.05 3.60 16 17.46 15 1300 BoxM 2.96 2.74 17 17.46 15 1250 Box" 2.89 4.09 18 17.46 15 1300 Box" 2.98 5.10 19 17.46 15 1250 Box" 2.84 N/A 20 250 15 1300 Box" 3.09 2.30 * Tube fumace Thermtec, Lenton Thermal Designs ** Box furnace C+M Inc., High Temperature Furnaces Bloomfield NJ 57 Table 3-1 continued Specimen Pressure Sintering Sintering Furnace Density Grain size label (MPa) time temperature type (g/cm’) (microns) (nrin) (0 C) 21 250 120 1300 Box'" 3.09 2.60 22 250 15 1200 Box“ * 2.92 1.99 23 250 30 1250 Box“ * 3.07 2.06 24 6.55 15 1200 Box‘" 2.47 4.44 25 6.55 30 1050 p-wave* 2.26 2.73 26 6.55 30 l 100 u-wave" 2.65 3.09 27 6.55 30 1 150 u-wave“ 2.95 3.85 28 6.55 30 1200 u-wave" 2.98 4.36 3 l 80 60 1200 Box“ * 3 .02 5 .05 32 80 60 1350 Box" * N/A 4.78 33 so 60 1 100 Box" 1.33 1.48 34 80 60 1250 Box‘MI 3.06 2.54 35 80 60,130 I 150,1300 Box“ "‘ 3.02 2.70 36 13.09 60 1100 Box** 2.08 2.87 37 19.64 60 1100 Box" 2.17 3.92 38 6.55 60 1100 Box'” 1.54 N/A 39 26.18 60 1200 Box* * 2.80 2.54 40 13.09 60 1200 Box“ * 2.59 2.67 41 6.55 60 1200 Box" * 2.43 2.40 42 13.09 60 1250 Box'MI 2.95 2.80 43 19.64 60 1250 Box** 2.99 3.31 * Microwave Furnace, Wavemat Cavity, Plymouth, Nfl; Power Supply, Sairem Vaux-en-Velin, France ** Box firmace C+M Inc., High Temperature Furnaces Bloomfield NJ 58 Table 3-1 continued Specimen Pressure Sintering Sintering Furnace Density Grain size label (MPa) time temperature type (g/cm’) (microns) (min) (°C) 44 26.18 60 1250 Box" 3.00 2.89 45 6.55 60 1350 Box” 2.96 4.59 46 13.09 60 1350 Box" 2.90 3.33 47 26.18 60 1350 Box" 2.98 N/A 48 13.09 60,130 . 1150,1300“ Box" 2.89 N/A 49 19.64 60,130 1 150,1300* Box" 2.92 N/A 50 26.18 60,130 1 150,1300“ Box" 2.94 4.46 " First run at 1150° C for 60 minutes and specimens came out badly cupped, second run at 1300° C for 130 minutes. ** Box furnace C+M Inc., High Temperature Furnaces Bloomfield NJ 59 98 ~- .é‘ 8 “ i 8 96 -- o E, -- X 2 C I - o __ 05 94 o - I Be .. I 92 ~- .. II-i'h 90rii‘rii:iriitit 0 100 200 300 400 500 600 700 Sintering Time (minutes) I 6.6 MPa 0 17.5 MPa 2 24 MPa Figure 3-1 Density as a function of sintering time and different uniaxial die pressures for hydroxyapatite sintered in the conventional furnaces at 1300 °C. The data was taken from a subset of the data presented in Table 3-1. 60 Figure 3-2 ‘0 on J l 'i" I ea.-- 5 o 1' I 394.. E I d) a: .. 3292 -- 90 . : —: i 4 : 5 10 15 20 25 Die Pressure (MPa) I Experimental data — Linear Fit Density as a function of die pressure. The data was taken fiom sintered hydroxyapatite specimens sintered at 1300° C for 15-120 minutes, which are a subset of the data presented in Table 3-1. Linear fit represents the least squares fit of the data to equation 3-1. Fit parameters are slope = 0.27% relative density per 1 MPa die pressure, y-intercept = 91% relative density and the coefficient of determination r2 = 0.997. 61 100 a .. II II II II I 90 -- . '3; so -- ' 5 .. I Q ,3 7o -.- - a i- '- 62 so -- - e\° .. 50 "ii- - 40 : i i i : i : i 4. : : : : i 1050 1100 1150' 1200 1250 1300 1350 1400 Sintering Temperature (C) Figure 3-3 Density as a function of sintering temperature for hydroxyapatite specimens sintered for 60 minutes. The data was taken from a subset of the data presented in Table 3-1. 62 -\ -- V I ”’ v E ' u x Q 4 h V . x e ,_ , v v g I v x I U) I t 2 -- I X E U) 1’ ‘ 0 : i : t : : 4 : 1000 1100 1200 1300 1400 Sintering Temperature (C) I 15 min 0 Microwave 30 min. X 45 min. V 60 min. Figure 3-4 Average grain size of sintered hydroxyapatite as a function of sintering temperature and time for both conventional and microwave sintering. The data was taken fi'om a subset of the data presented in Table 3-1. 63 relationship for microwave sintered hydroxyapatite samples is linear for the small sample size of 4 specimens (Figure 3-4). The grain size was decoupled from the density for sintered hydroxyapatite specimens with densities greater than 2.90 g/cm3 (Figure 3-5). 3.1.3 Etched surface morphology Uniaxially pressed sintered hydroxyapatite specimens were polished and thermally etched for the grain size measurements. The thermally etched surfaces of the uniaxially pressed sintered HAP specimens were composed of polygonal grain shapes (Figure 3-6 and Figure 3-7). Polished surfaces of the hydroxyapatite surfaces were also acid etched to delineate the grain boundaries and measure the grain size. Acid etching (HAP specimens numbers 1-5, 11) resulted in severe changes in the surface due to over etching (Figure 3- 8). Polished and thermally etched surfaces of the hydroxyapatite specimens cold isostatically pressed to 250 MPa (numbers 20-23) exhibit significantly different etched surface morphology. CIPed specimens sintered for 30 minutes or less at temperatures of 1300°C or less had average grain sizes of 1.99 um , 2.06 pm 2.3 um (Figure 3-9, Figure 3-10, Figure 3-11 respectively). Specimen number 21, sintered at 13000 C for 120 minutes, grew the familiar polygonal shaped grains with an average grain size of 2.60 microns (Figure 3-12). 3.1.4 X-ray diffraction The difli’action scan, using Cu Kerl radiation (A = 15.406 nm), was from 25 degrees to 36 degrees where the most intense peaks from HAP and the decomposition 64 .. I 7 -u— .. I A6 -- I g .. I a5 -- '- 3 " I ‘ I o4 .. . I .5 I (D .. .- 5 3 J- I ' g h - I I - I‘ . - r 2 -. fl . - I 1 _. 0 i . 4 . i . i 55 65 75 85 95 % Relative Density Figure 3-5 Grain size as a function of percent relative density for sintered hydroxyapatite. The data was taken from a subset of the data presented in Table 3-1. 65 Figure 3-6 Micrographs of polished and thermally etched surface of sintered hydroxyapatite specimen number 15 sintered at 13000 C for 45 minutes, average grain size = 3.60 microns. 66 Figure 3-7 V _.__ g; Micrograph of polished and thermally etched surface of sintered hydroxyapatite specimen number 45 sintered at 1350° C for 60 minutes, average grain size = 4.59 microns. 67 Figure 3-8 -. /‘ Tn“ ' 7" I ‘1'» r. (”1“ .; “if .33 . ~,~\. g} s; * «w‘ .. (a) (b) Micrographs of polished and acid etched surface of sintered hydroxyapatite (0.1 M HCl for 1 minute), (a) specimen number 3 sintered at 1300°C for 45 minutes, grain size =1 .98 microns, (b) specimen number 8 sintered at 14000 C degrees for 240 minutes, grain size = 7.44 microns. 68 a. ‘1 if} iii ifi if: 1:: E? Figure 3-9 Polished and thermally etched surfaces of cold isostatically pressed (250 MPa) and sintered hydroxyapatite specimen number 22 sintered at 12000 C for 15 minutes. Average grain size = 1.99 pm 69 Figure 3-10 Polished and thermally etched surfaces of cold isostatically pressed (250 MPa) and sintered hydroxyapatite specimen number 23 sintered at 1250° C for 30 minutes. Average grain size = 2.06 pm 70 <- 10 microns —> Figure 3-11 Polished and thermally etched surfaces of cold isostatically pressed (250 MPa) and sintered hydroxyapatite specimen #20 sintered at 13000 C for 15 minutes. Average grain size = 2.30 mm 71 <— 7.5 microns => Figure 3-12 Polished and thermally etched surfaces of cold isostatically pressed (250 MPa) and sintered hydroxyapatite specimen number 21 sintered at 1300° C for 120 minutes, Average grain size = 2.60 um. 72 products of hydroxyapatite exist. According to the Powder Diffraction File for Inorganic Phases [68] the three main peaks for hydroxyapatite are at 20 = 31.80 for {211}, 20 = 322° for {112} and at 20 = 330° for the {300}. The (Jr-tricalcium phosphate (ct-TCP) primary peak is located at 20 = 30.7°, the B—tricalcium phosphate (B-TCP), primary peak is at 20 = 31 .0° and the tetracalcium phosphate primary diffraction peak at 20 = 289° [68]. The major peaks in the x-ray difli'action patterns, for the as received hydroxyapatite powder (Raw HAP) and sintered hydroxyapatite specimens numbers 4-12, are those of hydroxyapatite (Figure 3-13 and 3-14). A faint peak at 20 ~ 29° for hydroxyapatite specimens (numbers 7, 8, 9, 11, and 12) is associated with tetracalcium phosphate. Tetracalcium phosphate is the minor product of the decomposition reaction by a factor of 1:2 compared to tri-calcium phosphate (Equation 1-2). Tri-calcium phosphate was not detected in the sintered hydroxyapatite specimens by x-ray difli‘action. A decrease of the lattice parameter co is noted in sintered hydroxyapatite specimens as the sintering temperature and sintering time increases (Table 3-2). The lattice parameter were calculated from the angles of the three major x-ray difl‘raetion peaks of the hydroxyapatite specimens (Figures 3-13 and 3-14) using Equation 2-2 and 2- 3. The lattice parameter CO was determined by averaging the lattice parameter c0 calculated fi'om the diffraction peak for the {211} plane at 20 = 31.8° and the dim-action peak for the {112} plane at 20 = 322°. The lattice parameter a(, was calculated using the peak for the {300} plane at 20 = 33.0°. Sintered hydroxyapatite specimens numbers 4-12 are composed of hydroxyapatite with no evidence of decomposition. 73 #1) . I i , i it—9—HAP .59 ..____. g #8 HAP o —- #7 HAP #6 HAP Oriitiiiiiiiiiii++iifiiifi‘ HAPpowder 25 27 29 31 33 35 2-theta (degrees) Figure 3-13 X-ray difl‘raction pattern (Cu K,ll radiation) for as received (Raw) and sintered hydroxyapatite (HAP). Specimens number 6 and 9 sintered at 13000 C for 240 and 300 minutes respectively. Specimens number 7 and 8 sintered at 14000 C for 120 and 240 minutes respectively. * Primary tetra-calcium phosphate diffraction peak. 74 Counts #12 HAP #11 HAP 25 27 29 31 2-theta (degrees) Figure 3-14 Additional x-ray difli'action patterns (Cu K“l radiation) for sintered hydroxyapatite. Specimen #11 and #12 sintered at 13000 C for 540 and 660 rrrinutes respectively. * Primary and secondary tetra-calcium phosphate difii’action peaks. 75 Table 3-2 Lattice parameter data for the as received hydroxyapatite powder (HAP I) and sintered hydroxyapatite, calculated using the 3 major hydroxyapatite peaks fi'om Figures 3-13 and 3-14 and Equations 2-2 and 2-3. Data fiom the handbook is included [69]. Specimen Sintering Sintering a0 c0 label time (min) temperature (°C) (nm) (nm) HAP I 0 --------------- 0.9389 06943 6 120 1300 0.9406 0.6892 9 300 1300 0.941 1 0.6903 10 420 1300 0.9406 0.6857 1 1 540 1300 0.9384 0.6877 12 660 1300 0.9400 0.6868 7 120 1400 0.9400 0.6860 8 240 1400 0.9400 0.6859 Average ------------------------- 0.9400 0.6882 PDF [69] File #9-432. -------------- 0.9418 0.6884 3.2 Indentation measurements Hardness and fiacture toughness of the sintered hydroxyapatite specimens were measured by the Vickers indentation method (Table 3-3). 3.2.1 Vickers hardness Vickers hardness of the sintered hydroxyapatite specimens prepared in this study are a strong function of the density (Figure 3-15). As a firnction of the average grain size the hardness values of the sintered hydroxyapatite specimens show no trends indicating that grain size effects on hardness are significantly weaker than the porosity effects on hardness, over the range of grain sizes and porosity values for the specimens in this study (Figure 3-16). There are no discernable patterns in the standard deviation (of the 7-10 76 Table 3-3 Density, grain size and the mechanical properties of hardness and toughness, measured by indentation technique, of sintered hydroxyapatite. The standard deviations (0), calculated for the multiple indentations per specimen, are included. Specimen %Relative Grain size Hardness O Fracture (J Fracture label density x10°m GPa Hardness toughness toughness GPa MPa(m”2) MPa(m"2) 1 91.4 1.68 3.543 0.274 1.125 0.266 2 91.3 1.65 3.333 0.248 1.233 0.238 3 91.2 1.98 2.792 0.234 1.466 0.231 4 91.5 2.07 3.758 0.432 1.017 0.294 5 91.4 2.14 3.774 0.515 1.006 0.187 6 92.5 2.32 3.326 0.281 1.209 0.204 7 96.6 5.65 4.816 0.551 0.794 0.259 8 97.0 7.44 4.840 0.189 0.479 0.037 9 94.0 5.93 4.145 0.245 0.715 0.140 10 93.6 5.93 3.671 0.150 0.833 0.354 1 1 94.4 6.73 4.197 0.206 0.732 0.104 12 94.2 6.80 4.428 0.351 0.701 0.144 13 96.4 4.37 3.948 0.660 1.017 0.230 14 95.1 3.70 3 .935 0.694 1.027 0.290 15 96.6 3.60 4.446 0.421 0.877 0.149 16 93.6 2.74 3.809 0.165 0.928 0.111 17 91.4 4.09 4.154 0.647 0.789 0.275 18 94.4 5.10 4.347 0.427 0.773 0.182 19 90.0 N/A 3.999 0.229 0.803 0.157 20 97.9 2.30 5.52 0.214 0.450 0.050 21 97.9 2.60 5.42 0.179 0.37 0.066 22 92.6 1.99 3.55 0.471 1.05 0.236 77 Table 3-3 continued Specimen %Relative Grain size Hardness 0 Fracture O Fracture label density xlof’m GPa Hardness toughness Toughness GPa MPa(m"2) MPa(m"2) 23 97.3 2.06 5.21 0.191 0.63 0.160 24 78.2 4.44 1.32 0.323 2.11 0.439 25 71.7 2.73 0.91 0.165 2.49 0.259 26 84.0 3.09 2.46 0.379 2.07 0.397 27 93.5 3.85 4.48 0.451 0.76 0.141 28 94.3 4.36 4.07 0.259 0.73 0.089 31 95.7 5.05 5.24 0.222 0.60 0.083 32 N/A 4.79 5.23 0.266 0.38 0.068 34 97.0 2.54 5.54 0.258 0.59 0.070 35 95.6 2.70 5.33 0.237 0.97 0.119 37 68.8 3.92 1.20 0.136 1.84 0.721 39 88.6 2.54 3.68 0.182 1.50 0.161 40 82.2 2.67 1.61 0.106 2.34 0.309 41 76.9 2.40 1.60 0.063 2.67 0.242 42 93.6 2.80 4.15 0.347 0.872 0.166 43 94.7 3.31 4.44 0.336 1.25 0.369 44 95.2 2.89 5.37 0.262 0.695 0.051 45 93.7 4.59 3.77 0.293 0.996 0.152 48 91.4 N/A 4.20 0.176 0.790 0.062 49 92.4 3.21 4.20 0.309 0.868 0.122 50 93.1 4.46 3.08 0.109 1.45 0.319 78 .h r l ' l I '- I Hardness (GPa) (a) I 1 I N L 1 l l l 65 75 85 95 % Relative Density 0 ‘- Figure 3-15 Vickers indentation Hardness values for sintered hydroxyapatite as a function of density. The data was taken from Table 3-3. 79 .. I 54 _ I- _. ' I I -- III ’64." I. III - I : CL I 0 .. If '- I ’ - 2 “ " s2" .. II I . I ‘14- . 0 i : i i i : i : i i =1 1 2 3 4 5 6 7 8 Grain Size (microns) Figure 3-16 Vickers Hardness plotted against the average grain size of the sintered hydroxyapatite. The data was taken fi‘om Table 3-3. 80 hardness measurements per specimen) as a function of grain size (Figures 3-17). 3.2.2 Indentation fracture toughness Fracture toughness of the sintered hydroxyapatite specimens were measured to be 0.4 - 0.6 MPa m”2 at ~ 97% relative density. The fracture toughness of the sintered hydroxyapatite specimens increases markedly, to 2.5-2.7 MPa mm, as the density decreases to between 72 and 77% relative density (Figure 3-18). A plot of the fracture toughness as a function of the grain size shows no trend over the grain size range of the specimens in this study (Figure 3-19). Standard deviation of the fiacture toughness measurements plotted against the specimen’s average grain size shows no trend (Figure 3- 20). The ratio between the indentation crack length (c) and the indentation diagonal (a), as a function of relative density, increases as the density increases for the sintered hydroxyapatite specimens in this study (Figure 3-21). The trend for the c/a ratio as a firnction of relative density shows that the cracks barely extend beyond the indent for percent relative density below 85%. 3.3 Hardness and microstructure comparisons with literature data All of the Vickers hardness data of the sintered hydroxyapatite specimens fiom this study (indentation loads of 0.3 kg, 0.5 kg and 1 kg), and selected Vickers hardness data from Lu et al. (indent load 0.05 kg) [74], and Best et al. (indent load 1 kg) [75] along with selected Knoop hardness data fi'om Wang et al. (0.2 kg load) [9] and Slosarczyk (no load specified) [11] all show substantial decreases in hardness as the porosity increases 81 A07-i- I E x '- 9.0.6-L .5 i I gas-- .- 8 dr- . |-- I “904* - a -— I ' 503-— : ' co - I 802» ' ' = ' ' 5 E ‘ I m :01-- I I -. I 0 a i . : . i 0 2 4 6 8 Grain Size (microns) Figure 3-17 Statistical standard deviation of the Vickers hardness of sintered hydroxyapatite plotted against the average grain size. The data was taken fiom Table 3-3. 82 w J 1 l I A O $2.5 -- o E .. O (I! O % 2 -- ‘ 7; o m .. ‘c’ §15 -- . . . .2 " g: ' C) -r- 5 1 .. 9 g .t O O ' LL0.5 -_ ' 2’ 0 i : : 4 i 65 75 85 95 % Relative Density Figure 3-18 Vickers indentation toughness as a fimction of density for sintered hydroxyapatite. The data was taken from Table 3-3. 83 A “P . $2.5 -- o E __ C to C E 2 -r . en 0 m -. 3 .§’1.5 "" . . . a " g . ' g 1-- ~ ‘ .. .. . o E "‘ o "O . . 0 Q LL05 __ . O O . ' ° 0 o 0 x i i L i u i 1 2 3 4 5 6 7 Grain Size (microns) Figure 3-19 Fracture toughness as a function of average grain size of sintered hydroxyapatite. The data was taken from Table 3-3. 84 0.8 1" g0] -_ ’ 2' a- E0.6 ,_ (U 0.. q- gas ~- C 8 *’ 0 £04 -- Q. a -- . , . '90-3 7" O . g g .. O o . o c . . . O 30.2 -> .0 . w " . .. .0 o 0 g o 01 "" . . . .. .c. o o 0 i l u i 0 2 4 6 8 Grain Size (Microns) Figure 3-20 Standard deviation of the fracture toughness plotted against the average grain size of the sintered hydroxyapatite. The data was taken fiom Table 3-3. 85 6 __ 5 __ .9 E4 "H- a -h- 33 -- I __ I 2 —- _ ' " r 1 a— I I I I I I 0 1 1' , i 4 l i - i . 1‘ 1 l 65 70 75 80 85 90 95 1 00 % Relative Density Figure 3-21 Ratio of the Vickers indentation crack length and Vickers indentation diagonal plotted against the % relative density, for all 410 indents on the specimens with known density measured in this study. 86 (Table 3-4). The selection criteria for the hardness data, fiom the literature, limited the data to those specimens proved by x-ray difiraction to contain no more than trace amounts of the decomposition phases of hydroxyapatite. The hardness data of sintered hydroxyapatite, from this study (Table 3-3) and the hardness data from the studies by Wang et al. [9], Slosarczyk et al. [11], Best et a1. [75], and Lu et al. [74] (Table 3-4), were linearized using the model proposed by Rice [57] (Equation 3-1) lnH=h1Ho-bP (3'1) where H is the measured hardness, H0 is the zero porosity hardness, and b is the slope of the plot of the logarithm of hardness (H) versus volume fraction porosity (P) (see Equation 1-8 in section 1.6). Linear regression analysis (Quattro pro) of the linearized data was used to calculate the zero porosity intercept HO and the b value (Table 3-4). The hardness data and the linear fit of the hardness data was then normalized by dividing the measured hardness by the zero porosity hardness (Ho). The data for Wang et al. and [9] Slosarczyk et al. [1 1] were linearized and normalized together due to the conformity of the two data sets. The normalized data and the normalized linear fit of the hardness data shows that the data from this study compares favorably to the data fi'om Best et a1. [75], and Wang et al. [9] Slosarczyk et a]. [11] as a function of the volume fiaction porosity (Figure 3-22). The hardness values from Lu et al. [74] decreases at a significantly greater rate as a fimction of the volume fraction porosity then the rest of the plotted data. Hydroxyapatite powder used in preparing the specimens for Wang et al. [9], and Slosarczyk et al. [1 1] was synthesized by a wet chemical-method. Lu et al. [74] prepared two separate hydroxyapatite powders using the wet chemical method by heat drying part 87 Table 3-4 The normalization and statistical parameters from the linear regression analysis of the hardness as a function of volume fraction porosity (Equation 3-1: In H = 1n Ho - bP) for this study, Wang et al. [9], Slosarczyk et al. [1 1], Best et al. [70], and Lu et al. [69]. Actual hardness data [9,11,69,70] from the literature is included. Linear fit parameters calculated from the hardness data in this study (Table 3-3) b = 6.03 r2 = 0.907 Standard error" = 0.131 Calculated zero porosity hardness Ho = 6.00:1: 0.79 Gpa Wang et al. [9] and Best et al. [70] Lu et al. [69] Slosarczyk et al. [11] Data fiom tables Data fi'om graphs Data from graphs b = 7.38 b = 4.89 b = 19.4 r2 = 0.892 r2 = 0.947 r2 = 0.969 Standard error” = 0.379 Standard error" = 0.244 Standard error" = 0.125 Ho = 5.18 :t 1.96 GPa Ho = 552.9 i 135 Ho = 1218.4 :4: 152 Volume Knoop Volume Vickers Volume Vickers fiaction hardness fi'action hardness fiaction hardness porosity (P) Hk (GPa) porosity (P) H, Unitless porosity (P) IL Unitless 0.40 0.449 0.52 50 0.12 125 0.39 0.155 0.50 50 0.11 128 0.38 0.189 0.49 50 0.05 560 0.35 0.298 0.48 51 0.05 580 0.31 0.85 0.18 145 0.04 605 0.31 0.849 0.16 350 0.04 648 0.24 0.928 0.1 l 205 0.03 660 ** Standard error of the normalized hardness (H/Ho) 88 Table 3-4 continued Wang et al. [W] and Best et al. [B] Lu et al. [L] Slosarczyk et a]. [S] Data from tables Data from graphs Data fi'om graphs Volume Knoop Volume Vickers Volume Vickers fiaction hardness fraction hardness fiaction hardness porosity (P) Hk (GPa) porosity (P) PL Unitless porosity (P) H, Unitless 0.23 0.963 0.09 370 0.03 680 0.22 1.025 0.06 425 0.02 758 0.08 2.7* 0.05 410 0.01" 805" 0.02 4.0* 0.03 410 V 0.01 4.6* 4 0.03 450 \ 0.02 700 0.02 670 ‘ k 0.01 540 \\ * Hardness data fi'om Sosarczyk et al. [S] it 89 Trace of TCP in x-ray difii‘action spectra. 10 3 0 C E m I .3 b= 4.9 E 0 m C 0.01 I i i ; 1L ; 1' ; i ' rf j. 0 0.1 0.2 0.3 0.4 0.5 Volume Fraction Porosity (P) I Present Study. [9,11] 2 [70] V [69] Figure 3-22 Log relative indentation Hardness values for sintered hydroxyapatite plotted against the volume fraction porosity along with the solutions for the linear log relative hardness as a fimction of volume fiaction porosity (Equation 1- 8, Section 1.6). Data was taken from Table 3-3 for the present study and Table 3-4 for the literature data. 90 of the mixture of HAP and water, and freeze drying the rest. Hardness data from specimens produced from both powders by Lu et al. [74] were lumped together for analysis (Table 3-4). Three commercial hydroxyapatite powders, used by Best et al. [75], were characterized by x-ray diffi'action, infrared spectroscopy, inductively coupled plasma spectroscopy, surface area analysis, particle size analysis and scanning electron microscopy, before being sintered. The three commercial powders were found to have significantly different morphologies which influenced the sintered density and grain size of the sintered specimens [75]. Hardness data fi'om specimens sintered from all three commercial powders were included in this analysis (Table 3-4). The resultant microstructures of the sintered hydroxyapatite specimens compiled in table 3-4 included grain sizes of 2-8 pm this study, 5-10 pm [11,74], to 90 um [9]. Grain sizes of the final sintered specimens were not specified by Best et al. [75]. The data fi'om this study is in good agreement with the four different studies [9,11,74,75]. Indentation hardness of sintered hydroxyapatite decreases exponentially as porosity increases, as modeled by equation 1-8, regardless of the origin and morphology of the starting powder. 3.4 Fracture toughness and microstructure comparisons with literature data Fracture toughness studies of sintered hydroxyapatite show a decrease in the fracture toughness values as the porosity increases when measured by three point bend [54], and the indentation method [56] (Table 3-5). The fi'acture toughness of sintered hydroxyapatite specimens, in this study, increased in fracture toughness as the porosity increases (Figure 3-23). The fi'acture toughness of the whisker reinforced sintered 91 Table 3-5 The normalization and statistical parameters fi'om the linear regression analysis of toughness as a function of volume fi'action porosity (Equation 3-2: In K, = 1n Koic - b P) for this study, DeWith et al. [54] and Suchanek et al. [56]. Fracture toughness data fi'om DeWith et al. [54] and Suchanek et al. [56] are included. Linear fit parameters from the fiacture toughness data in this study (Table 3-1) b=-5.51 r2 = 0.657 Standard error" = 0.270 Calculated zero porosity fracture toughness K,Co = 0.61 :t 0.16 MPa m"2 DeWith et al. [54] Suchanek et al. [56] Data from graph Data from graph b = 2.786 b = 2.812 r2 = 0.630 r'2 = 0.786 Standard error” = 0.202 Standard error" = 0.116 K.Co = 1.14 :1: 0.23 MPa rn“2 Kn, =1.63 :1: 0.19 MPa :11"2 Volume fiaction Fracture Toughness Volume fraction Fracture Toughness Porosity (P) (MPa mm) Porosity 0’) (MPa mm) 0.27 0.65 0.29 0.78 0.27 0.6 0.16 1.00 0.22 0.64 0.1 l 1.10 0.22 0.35 0.10 1.05 0.18 0.66 0.08 1.36 0.18 0.81 0.08 1.41 0.1 0.93 0.05 1.32 0.1 0.88 0.04 1.54 0.06 0.98 0.03 1.34 0.06 0.93 0.02 2.00 ** Standard error of the normalized fracture toughness (K,C /K.Co) Table 3-5 continued DeWith et al. [54] Suchanek et al. [56] Data fiom graph Data from graph Volume fiaction Fracture Toughness Volume fraction Fracture Toughness Porosity (1’) (MPa mm) Porosity (1’) (MPa mm) 0.03 1.08 0.02 1.45 0.03 1.01 0.02 1.65 ‘ \V' 0.01 1.48 0.01 1.44 \ 0.01 1.82 k 0.01 1.72 93 0) J l g25 "’ - < . -. I E __ I d.“ 2..L . - - 3 +3 I 3 O m1.5" C c Ill 0) 3 1- O O ,_ . x o a) X X i 50.5 ‘6 . x i! u. 0 . i 4 i 4 i . i . i t i t i 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 Volume Fraction Porosity (P) I Present Study 0 [56] x [54] Figure 3-23 Fracture toughness values for sintered hydroxyapatite plotted against the volume fraction porosity. Fracture toughness values were taken from Table 3-3 (present study) and Table 3-5 for the data from Suchanek et al. [56] and DeWith et al. [54]. 94 hydroxyapatite, prepared by Suchanek et a1 [56] is greater than the fiacture toughness, as a function of the porosity, of the over the non-reinforced specimens prepared by DeWith et al. [54] (Figure 3-23). The fi'acture toughness data from DeWith et al. [54], Suchanek et a1. [56] and this study were linearlized and normalized in the same way as the hardness data (section 3.3). ant=anw-bP (3-2) where K,c is the measured fracture toughness, Kid, is the zero porosity fracture toughness, and b is the slope of the plot of the logarithm of fiacture toughness ( Kt) versus volume fiaction porosity (P) (see Equation 1-8 in section 1.6). The data (Table 3-5) from DeWith et al.[54] and Suchanek et al [56] have the same functional relationship to the volume fraction porosity, as determined by the b value (Figure 3-24, and Figure 3-25). The normalized and linearized data from this study was. not linear and the calculated b value was negative, deviating from the MSA model (Figure 3-24, and Figure 3-25). This study and the study by DeWith et al. [54] used commercially available powders (CERAC Inc., this study, and Merck A. G. [54]). Both studies have similar microstructures. The sintered grain sizes are 1-6 pm for DeWith [54] and 2-8 pm for this study. Porosity values were up to 0.30 volume fiaction porosity for both studies. Traces of decomposition products were detected in specimens sintered at temperatures greater than 11500 C (relative densities > 85%) for DeWith et al. [54]. The indentation fiacture toughness method used by Suchanek et a1. [56], is similar to the fracture toughness measurement technique used for this study, however, the microstructures difler due to the whisker reinforcement in the sintered hydroxyapatite specimens tested by Suchanek et al. [56]. 95 1° b = -5.5 m 3 I c - I g, I D O '— d) 5 1 E x LL .8 x % b=2.8 o: 0.1. : +r::t+::::‘fi:i 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 Volume Fraction Porosity (P) I Present Study 0 [56] x [54] Figure 3-24 Relative fracture toughness values for sintered hydroxyapatite as a function of volume fraction porosity and the calculated linear fit. Linear fit parameters taken from Table 3-5 for the data from Table 3-3 (present study) and the the toughness data in Table 3-5 fi'om Suchanek et al. [56] and DeWith et al. [54]. 96 .s O b = -5.5 Relative Fracture Toughness b=2.8 0.1 . a i : : 4. i r : : : : : 0 0.025 0.05 0.075 0.1 0.125 0.15 Volume Fraction Porosity (P) I Present Study 0 [56] x [54] Figure 3-25 A section of lower volume fi'action porosity of the relative fiacture toughness values for sintered hydroxyapatite as a function of volume fi'action porosity and the calculated linear fit. Linear fit parameters taken from Table 3-5 for the data fi'om Table 3-3 (present study) and the fi‘acture toughness data in Table 3-5 from Suchanek et al. [56] and DeWith et aL [54]. 97 The studies by DeWith et a1. [54] and Suchanek [56] are linked by the fact that the fracture toughness measurement techniques and sintered hydroxyapatite microstructure were vastly different yet the fimctional relationship between fracture toughness and volume fi'action porosity is nearly identical. Comparisons between the three studies leads to a three part argument, in which no conclusion can easily be drawn; 1) the starting HAP powders and sintered HAP microstructure are similar in this study and the study by DeWith et al. [54], yet the nearly fully dense (95-97% relative density) fracture toughness and the functional relationship between fi'acture toughness and porosity are vastly difi‘erent. 2) the studies by DeWith et al. [54] and Suchanek et al. [56] are linked by a nearly identical b value calculated fi'om the fiacture toughness versus volume fraction porosity data (Equation 3-2) even with very different measurement methods and specimen microstructures. 3) the indentation fiacture toughness techniques are shared by this study and Suchanek et al. [56], however the measured fimctional relationship between fiacture toughness and porosity are incongruent. 3.5 Further comment on the fracture toughness of porous hydroxyapatite The significant and steady increase in the fiacture toughness of the sintered hydroxyapatite specimens in this study as the porosity of hydroxyapatite specimens increases (Figure 3-18) requires further comment. The lack of evidence of secondary phase formation, from x-ray difliaction (section 3.1.4), the large sample population (42), and the increase of the fracture toughness as the density decreases, forces a focus on the fi'acture toughness as a firnction of porosity for the sintered hydroxyapatite specimens in this study. 98 The relationship among the microstructure of sintered hydroxyapatite, the indentation fracture toughness measurement and the measured fi'acture toughness of sintered hydroxyapatite consists of three major elements: 1) Applications of the indentation toughness equation (Equation 1-6) to porous commie materials. 2) The formation and grth of indentation cracks in porous hydroxyapatite specimens 3) Toughening mechanisms that are unaffected by the grain size, over the grain size range of the specimens in this study, and are increasingly effective for hydroxyapatite specimens in which the processing parameters produced higher porosities. The fracture toughness of near firll density sintered hydroxyapatite, reported in the literature, is between 0.8-1.2 MPa m"2 [54,56,76], and is comparable to the average of the +90% dense specimens in this study (ch = 0.77 MPa mm). 3.5.1 Indentation toughness model and porosity The stress field of the test specimen under indentation loading is a function of the elastic modulus and Poisson’s ratio of the test specimen, and the load and geometry of the indenter [3 8]. Indentation fracture toughness is calculated using the indentation stress field and the resultant indentation crack length (Figure 1-1 [42]) of the indented specimen [3 8] . The indentation toughness model, derived by fi'acture mechanics fi'om the indentation stress and the indentation cracks [3 8], has successfully determined the fiacture toughness for a wide variety of ceramic materials as a function of the hardness, elastic modulus and the indentation crack length [39,40,41] (Equation 33 [40]). 99 K = 4(3)“? .3. (3-3) 1c H 63/2 V where C is the empirically determined constant (C = 0.016 [40]), E is the elastic modulus, HV is the Vickers hardness determined from the size of the indent (Hv = P/2a2, Equation 2-5), P is the indentation load, and c is the indentation crack length (fi'om Figure 1-1, section 1.5.1). Equation 3—3 was used to calculate Klc for each indent in this study. The average and standard deviation of the specimen’s K“, were then determined from the 7-10 indents/specimen. The elastic modulus of ceramics decreases as the porosity increases [35,54,58,60]. The elastic modulus/porosity data has been fit to the MSA model (ln(E) = ln(Eo) - bEPp, Equation 1-8), where E0 is the fully dense elastic modulus, bE is the material dependent constant and PI. is the volume fraction porosity [57] (Section 1.6). Poisson’s ratio for sintered hydroxyapatite also decreases as the porosity increases [53,54]. The porosity dependence of Poisson’s ratio is much weaker, as determined by the b values, than the porosity dependence on elastic modulus, for many ceramics (Table 3-6). The fracture toughness values of the hydroxyapatite specimens in this study (Tables 3-3, 3-5, Figures 3-18, 3-19, 3-24, and 3-25) were calculated by equation 3-3 with fixed elastic modulus (E = 115 GPa [56]). The functional dependence of hardness, as measured by the size and load of the indent, on the volume fi'action porosity of sintered hydroxyapatite has been measured in this study and fit to the MSA model [57] (ln(H,) = ln(Hvo) - bH PP Section 3.3). The relationship between hardness and density indicates deformation processes behave as 100 Table 3-6 Porosity dependence of the elastic modulus compared to the porosity dependence of Poisson’s ratio for experimental data fit to the MSA model (lnE = ln E0 - bEP, lnv = in v0 - va, Equation 1-8) by Rice [58]. Ceramic P range (%) bE bv vo Reference A1203 15-55 2.8 0 0.27 Coble and Kingery from [58] B60 0-8 4.0 0 0.155 Petrak et al. from [58] HAP 3-27 2.9 0.7 0.28 [54] fit in this study AIN 0-20 5.5 0.9 0.24 Hunter et al. from [58] Hm 5-15 5.2 0.6 0.26 Desmaison-Brut et al. from [531 TaN 1-15 4.6 0.6 0.25 “ ” TIN 1-22 3.0 0.2 0.22 “ ” ZrN 2-8 3.3 0.8 0.26 “ ” Graphite 13-32 3.5 1 0.31 Cost et al. from [58] 101 expected by the MSA model [57] for sintered hydroxyapatite. Inserting the MSA functions (modifications of Equation 1-7) for the elastic modulus and Vickers hardness in equation 3-3 results in equation 3-4. -bz Pr P E09 1/2 K,c(E(PP), H(Pp)) = {(H e-bHP,) 03/2 (3'4) 0 where E0 and HW are the fiilly dense values of elastic modulus and hardness, bE and bH are the material dependent constants (MSA fit [57]) for the elastic modulus and Vickers hardness respectively, and Pp is the volume fiaction porosity. The material dependent constants bE and bH are obtained by fitting the modulus-porosity and the hardness-porosity data to the MSA model (Equation 1-8). Equation 3-4 is simplified by combining the exponents (Equation 3-5). (bH‘bIE). _.__ , E , P K,C(E(PP),H(P,,))=e 2 ”470123,: (3-5) 0 The efl‘ect of porosity on the elastic modulus/Vickers hardness ratio may be written as (Eu/Hv0)mexP(1/2(bn - bP))- Rice [5 8], applying existing literature data to the MSA model, compiled decay constants for the elastic modulus and hardness for a large variety of porous ceramics (Table 3-7). The effect of the porosity dependence on the hardness and elastic modulus (as related by bE and b“), to Equation 3-5 with (P/cm) held constant, ranges from a factor 102 Table 3-7 Material dependent coefficients (bH and b5) fi'om the MSA model [5 7] for A1203, B4C, AIN and SiO2 ceramics calculated by Rice [58]. The HAP (hydroxyapatite) elastic modulus attenuation coefiicient was calculated for this study using DeWith et al.’s data [54]. The hardness attenuation coeflicient for HAP has been determined previously in this study (Table 3- 4). Coeflicient A1203 B4C AIN SiO2 HAP bE 2.6 4.0 5.5 4.3 2.93“ bH 6.3 5.3 7.7 3.2 6.03" Ab=bH-bE 3.7 1.3. 2.2 -1.1 3.1 [Reference] Wu & Rice Wu & Rice Boch et al. Park & *[54] Hench from [58] from [58] fi‘om [58] from [58] “[this study] 103 2.10 for alumina (Ab = 3.7) at 0.4 volume fraction porosity to a factor of 0.80 for S102 (Ab = -1.1) at volume fraction porosity of 0.4 (Figure 3-26). The E/H coefficient for hydroxyapatite = 1.86 (Ab = 3.1) at 0.40 volume percent porosity, was calculated by the hardness decay constant fi'om the data in this study, bH = 6.03 (Table 3-4), and the elastic modulus decay constant b5 = 2.93 calculated in this study from the sintered hydroxyapatite data fiom DeWith et a1 [54] (Table 3-7, Figure 3-26). 3.5.1.1 Effect of E(P) on the fracture toughness of sintered hydroxyapatite A decrease in hardness with increasing porosity was observed for the sintered hydroxyapatite specimens in this study. The change in the Vickers hardness as porosity increases is included in the fracture toughness calculation, due to the hardness and fracture toughness values being calculated for each individual indent by Equation 3-3. Adding the elastic modulus-porosity function (E = E0 e"b P” ) with decay constant bE= 2. 93 determined by DeWith et al. [54] to equation 33 results in Equation 3-6. P K,C(E(PP)) = 2 ’ Nil—”a 1v)-——"2 c -—2—,2 (3-6) Recalculation of the fracture toughness of the sintered hydroxyapatite specimens in this study, using equation 3-6 and bE = 2.93 [54], decreases the fracture toughness relative to the fracture toughness calculated with equation 3-3. The corrected fracture toughness values, determined by K,2(E(Pp)) (Equation 3-6), for the sintered hydroxyapatite 104 N I l z Alumina o I O ~r — é HAP 351.5 -- _____ E”, :7 4__ AIN o d o — m ,,,,,,, 9.’ 1 B40 :0 (:7 - to - Silica 0.5 i : ‘fi : : : : 0 0.1 0.2 0.3 0.4 Volume Fraction Porosity Figure 3-26 (B "2 as a fimction of the volume fiaction porosities normalized by (ED/Ho)”, the firlly dense values of the elastic modulus and hardness respectively. The decay constants bH and b; are compiled by Rice [58] for A1203, AIN, B4C, and SiO2 (Table 3-6). The b" for HAP was determined in this study, while b5 for HAP was calculated in this study using the MSA model and DeWith et al.’s data [54] (Table 3-6). 105 specimens in this study more than double as the volume fraction porosity increases from 0.05 to 0.20 (Figure 3-27). In addition to the efl‘ect of porosity the elastic modulus can be reduced by microcracking between grains in a polycrystalline ceramic caused by thermal expansion anisotropy [36,64,77]. The interactions between porosity and microcracking, and microcracking and fracture toughness are very complex and will be discussed among the possible toughening mechanisms in section 3.5.3. 3.5.1.2 Indentation fracture toughness of porous YBa2Cu30.,_,l Tancret et al. [78] measured the mechanical properties of porous YBa2Cu3O7,,2 high temperature superconductors. The high temperature superconducting specimens were prepared by cold isostatic pressing and pressureless sintering of commercial YBa2Cu3O7_x powders [78] . The elastic modulus (E), measured by three-point bending under quasi-static conditions, the fi'acture toughness (K22), measured by single edged notched beam (SENB), the Vickers hardness (Hv), and the Vickers indentation crack parameter (P/c’”) were recorded for specimen porosities between 5 and 31% (0.05-0.31 volume fraction porosity) and average grain sizes between 3 anle um (Table 3-8) [78]. The indentation crack parameter (P/cm) decreases as porosity increases for the YBa2Cu3OM [78] (Figure 3-28). The indentation crack parameter (P/cm) for the sintered hydroxyapatite in this study is parabolic in shape peaking at ~ 1.7 MPa m"2 at 0.18 volume fiaction porosity and falling at volume fraction porosities > 0.25 (Figure 3-28). In this study, using Tancret et al.’s YBa2Cu3O-,,, data [78], the porosity dependence of the mechanical properties E, K1c , HV and We”2 were fit to the MSA model [57] (Table 3-8). 106 2.5 —- ...L 01 l I ghness (MPa m"1/2) 0 0 p 01 l I Fracture Tou 0 1 t : : : 1 : i : i : : 1‘ i 0 0.05 ‘0.1 0.15 0.2 0.25 0.3 0.35 Volume Fraction Porosity Figure 3-27 Calculated fiacture toughness of the hydroxyapatite specimens in this study using the elastic modulus as function of volume fi'action porosity (K.c (E(P))} (Equation 36, material dependent constant bl3 = 2.93 [54]). 107 Table 3-8 Experimental data for YBa2Cu3O7,x from Tancret et al. [78] and the fit, for this study, to the MSA model ln(A) = ln(A0)- bAPP, where A represents the property (E, Hv, etc.), A0 = the property at zero porosity (y-intercept for the linear regression), bA is the slope of the log-linear relation, and Pp is the volume fiaetion porosity. Included for comparison are the fracture toughness values calculated, for each porosity, using equation 3-3 {Kic = C(E/H)”2 P cm}, for this study. Column 6 contains the fiacture toughness calculated with constant elastic modulus (E =120 GPa), {K,2(Indent, constant E)}. Fracture toughness values in column 7 were calculated using the measured specimen elastic modulus [78], in equation 3-3 {chandenfl}. PP E Hv P/c”2 K22 K,c*(Indent, K2, " Volume [78] [78] [78] [78] constant B) (Indent) traction (SENB) porosity (GPa) (GPa) (MPa mm) (MPa mm) (MPa mm) (MPa mm) 0.05 97.2 3.95 15.85 N/A 1.40 1.26 0.06 81.1 4.08 18.03 1.73 1.57 1.29 0.07 62.6 4.09 13.35 1.58 1.16 0.84 0.15 51.4 N/A N/A N/A N/A N/A 0.16 74.4 2.91 12.75 1.09 1.31 1.03 0.25 32 1.09 N/A 0.7 3 N/A N/A 0.29 22 0.64 7.65 0.43 1.68 0.72 0.31 19 0.5 5.05 0.37 1.25 0.50 MSA Fit of the above properties A0 120 7.08 20.5 2.54 1.34 1.35 b 5.57 7.98 3.89 5.84 -0.180 2.65 r2 0.898 0.953 0.898 0.963 0.023 0.734 N/A- data not available from Tancret et al. [78] * Calculated with Equation 33 setting E = 120 GPa, Hv from column 3 and P/c”2 from column 4. “Calculated with equation 3-3 using E fi'om column 2, Hv from colurrm 3 and P/c”2 fi'om column 4. 108 N 01 N O R? I - :3 I < E15 10 I I g d [I I :10 -- 2‘2 + '- O E. 5J- :- o 0 i a 4. . i 0 ‘0.1 0.2 0.3 Volume Fraction Porosity I This Study —— [78] Figure 3-28 Crack parameter P/cm as a function of the volume fraction porosity for hydroxyapatite from this study and for YBa2Cu3O,,, from Tancret et al. [78]. The line through the YBa2Cu3O7_,2 data is from the MSA fit (lnP/c”2 = ln(P/cmo)- bP P2,) of the crack parameter P/e”2 [78]. 109 From Equation 3-3, the indentation fracture toughness was calculated using the elastic modulus, Vickers hardness, and P/c”2 for each YBa2Cu2O7_x specimen. The resulting indentation fracture toughness values fit to the MSA model exhibited significantly smaller material dependent constant (bk = 2.65) than the bk value (5.84) determined from the MSA fit of the fi'acture toughness values determined by SENB [78] (Table 3-8, Figure 3-29). The fracture toughness values of the YBa.2Cu3O7_x specimens calculated by equation 3-3 with a constant elastic modulus (E, = 120 GPa), along with each specimen’s Vickers hardness and crack parameter (P/cm) are unvarying over the entire porosity range (Table 3-8, Figure 3-29). YBa2Cu3O7,x high temperature superconducting ceramic has a orthorhombic crystal structure [79]. YBa2Cu3O7,x is similar in the mechanical properties of the elastic modulus and hardness as hydroxyapatite (Table 3-9). 3.5.1.3 Porosity dependence of the empirical constant C The difference between the fi‘acture toughness of YBa2Cu3O.,,x determined by SENB and by the indentation method suggests the correction factor, C, may be a function of the volume fiaction porosity. The constant C had been determined by comparing indentation crack patterns of a number of ‘er11 behaved” reference ceramics [40] to fiacture toughness values determined by conventional fiaeture toughness techniques such as double torsion (DT) [39,40], and double cantilever beam (DCB) [40]. Using the MSA model for the SENB fi'acture toughness data to serve as the correct fiacture toughness values, as a function of porosity, and the MSA model fit (ln(Kk) = ln(Kmo) - bx PP, Equation 1-8) for the indentation fiacture toughness values, as a fiinction of porosity, a 110 ._L o LllllLlJ A 1-1 rrrrrrl I '1‘ I III ghness (mPa m"1/2) Fracture Tou 0.1 A. 1 4. i 1 : 4 4 4 : : ‘r i 0 . 0.1 0.2 0.3 Volume Fraction Porosity I SENB o Indent x ConstantE Figure 3-29 Fracture toughness YBa2Cu3O7_x of as a function of the volume fraction porosity determined by SENB [78] (Table 3-7), Indentation method Kk = C (Ell-I)"2 P/c”2 (Equation 3-3) with the data from Table 3-7, and the indentation method (Equation 3-3) with constant elastic modulus (E = 114 GPa) fiom Tancret et al. [78]. 111 Table 3-9 Comparison of the zero porosity mechanical properties of YBa2Cu3O7,x and HAP (C310(PO4)6(OH)2)- Material E0 Hv0 K120 YBa2Cu3OM“ 114 GPa [78] 5.0 GPa [78] 2.4 MPa In"2 [78] MSA Fit" 120 GPa [78] 7.1 GPa [78] 2.5 MPa m"2 [78] Ca.,(Po,),(0H)2 115 GPa [54] 6.0 GPa [This study] 1.1 MPa m"2 [54] 117 GPa [80] 6.4 GPa [22] 1.2 MPa m“2 [76] * YBa2Cu2O2,x data fit by Tancret et al. [78] using power law functions: E = 150(1'Pp)In ; Hv0 = H\,(,(l-P,,)m and K2, = K,,,,(l-P,,)m where PP is the porosity, and m is a microstructure dependent exponent. ** YBazCu3O-M data [78] fit to the MSA model (ln(A) = ln(A0)- bAPP), for this study (fiom Table 3-8) 112 correction factor as a function of the volume fraction porosity C(Pp) can be calculated for porous YBa2Cu3O7_x [78], (Equation 3-7). K,CO(SENB)e“"‘P” _ 2,2,2 K,CO(SENB) K,co(1ndent)e’b"P” ’ e K,CO(Indent) ((102): (3'7) where the difference between the material dependent parameters bK (Indent) and bK (SENB) are from the MSA model fit (ln(ch) = ln(cho) - bK Pp, Equation 1-8) of the YBa2Cu3O7_,2 data, AbK = bK (Indent) - bK (SENB), K220(SENB) is the zero porosity fracture toughness determined from the SENB data fit to the MSA model, Kmandent) is the zero porosity fi'acture toughness determined fi'om the indentation fiaeture toughness data fit to the MSA model. For the data fiom Tancret et al. [78] (Table 3-8) the correction equation C(Pp) = 1.88 e '3‘”. Adding C(PP) to the indentation fi‘acture toughness equation (Equation 3-3) relates the calculated fracture toughness from the indentation method to the fiaeture toughness determined by SENB and results in (Equation 3-8). E P K1. = 4 (Pix-12:6)”2 :372— (3-8) Application of C(PP) to the indentation fracture toughness of sintered hydroxyapatite (K,C{E(PP)}) (Equation 3-6) can be used as a first approximation to correct the indentation fiacture toughness method to porous ceramics (Equation 3-9). 113 l —-2 . 1 P K. 8(1),. 1. E112.» = 4(P.)E.e (gem 2“ 13-91 The fi'acture toughness correction (K,2{C(PP),E(P,,)}) of the sintered hydroxyapatite specimens in this study further lessens the porosity dependence of the fiacture toughness (Figure 3-30), as compared to the fiacture toughness calculated by equation 3-6 (K..{B(P.>}> (Figure 3-27). 3.5.2 Examples of deviations from the MSA model for fracture toughness From the 21 examples found in the literature in which the fiacture toughness of porous ceramics were measured only 4 show the fi'acture toughness increasing slightly or remaining unchanged as the porosity increases (Table 3-10). The normalized fi’acture toughness parameter data fiom Ricote et al. [48] and Case et al. [81] (Figure 3-31) depends less on the specimen porosity than the sintered hydroxyapatite specimens in this study (Figure 3-32). The critical volume fraction porosity (P,) for which physical properties drop precipitously is seen for Gd2O3 at Pc ~ 0.30 [81]. At P ~ 0.25 the fiacture toughness parameter of the sintered hydroxyapatite specimens drops off precipitously and appears to indicate the critical volume fraction for sintered hydroxyapatite. 3.5.3 Possible toughening mechanisms in sintered hydroxyapatite Fracture toughness is a measure of the energy needed to extend a crack [82]. Crack deflection by pores may increase the fiacture toughness as the porosity increases 114 2.5 '1- ‘3 .. F 2 -— . E 0 E v- . . . é o o . . «01.5 -- ’ . 3 o 0.0 ’ E -I- . . . 32 O E 1 ‘1' ‘. . . g .. .0 0 g 0 5.0.5 -- 0 4 : : : i : 'r 1r 4. : .L i : : 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 Volume Fraction Porosity Figure 3-30 The fi-acture toughness of the sintered hydroxyapatite specimens in this study calculated with C(PP) (Equation 3-8). 115 Table 3-10 Fracture toughness/porosity trends fit to the MSA model (ln(Kbo) = 1n(K,,,)- ber) fi'om the literature data for a variety of porous ceramics. The processing method (prc) and fi'acture toughness test method (meth.) are included. Ceramic (prc*) Porosity range % bk (meth.**) Reference A1203 (HP) 240 2.2 (DCB) Wu and Rice from [53,58] A1203 (PC) 4-42 3.4 (NB) Lam et al. fi‘om [53] A1203 (HP) 0-9 2 (WOF) Cappola and Bradt fi'om [53] A1203 (S) 0.44 3.4 (NB) Pabst from [53] A1203 (HP/S) 2-46 4.2 (NB) Claussen et a1. fiom [53] A1203 (S) 5-50 2.5 (N/A) Evans and Tappin fi'om [53] A1203 (HP) 2-14 no fit (DCB) [53] MgO (HP) 5-25 no fit (DCB) [53] Gd203 (CIP/S) 347 no fit (NB) [81] HAP (CIP/S) 3-27 2.79 (TPB) [54] fit in this study HAP (HP,S) 1-29 2.81 (Ind) [56] fit in this study HAP (CIP,S) 2-31 no fit (Ind) [this study] PZT (S) 2-15 2.4 (DT) Biswas and Fulrath fiom [53] PCT (N/A) 4-28 no fit (Ind) [48] YBa2Cu3O7,x 5-31 5.8 (SENB) [78] fit in this study YBa2Cu3O7,x 5-31 2.7 (Ind) [78] fit in this study SiC (8) 1-7 1.8 (N/A) Seshardi et al. fi'om [58] B4C (HP) 0-15 2.3 (DCB) Wu and Rice fiom [58] 34C (HP) 0-15 3.4 (DT) Hollenberg & Walther fi'om [53] Si3N4 (HP/S) 0-10 5.2 (Ind) Mukhoyadhyay et al. fiom [53] Si2N4 (HP/S) 0-50 2.4 (N/A) Rice et al. fi'om [53] *HP-hot pressed, PC-pressure cast, S-sintered, /S-and sintered, CIP-cold isostatic pressing, HP,S-hot pressed specimens plus sintered specimens, N/A-data not available. “DCB-double cantilever beam, NB-notch beam, WOF-work of fi'acture, TPB-three point bend, Ind-indentation crack measurement, DT-double torsion, SENB-single edge notched beam, N/A-data not available. 116 E». a) E S to CL 2 0 5 [48] «3 J 0 U- 81 3 . l 1 .5 E . I O . 0 Z 0 i ‘ : l : i 1 1 . l 0 0.1 0.2 0.3 0.4 0.5 Volume Fraction Porosity Figure 3-31 Norrmlized fracture toughness parameters (fracture toughness for Ricote [48], and fiacture energy for Case et al. [81]) studies in which the fracture toughness parameters increased or decreased slightly, over a significant volmne fraction porosity range. Normalization is done by dividing the fiacture toughness by the zero porosity fiacture toughness as determined by the intercept of the linear fit. The linear fit (included in plot) calculated for the entire range of Ricote et al’s data [48]. The linear fit calculated for the Case et al. [81] data was done for specimens with < 0.30 volume fraction porosity. The ceramics in question are calcium modified lead titanate for Ricote [48] and gadolinia for Case et al. [81]. 117 E a) E . E I as o. 9 I 3 6 I 9 LL 8 I I! '6 I E o .- Z O : i : l # i 1. i F i 0 0.1 0.2 0.3 0.4 0.5 Volume Fraction Porosity Figure 3-32 Normalized fracture toughness for sintered hydroxyapatite specimens in this study. Normalization is done by dividing the fiacture toughness by the zero porosity fracture toughness as determined by the intercept of the linear fit. The linear fit (included in plot) was calculated for hydroxyapatite specimens < 0.20 volume fi'action porosity. 118 [83,84]. For a specific crack-pore geometry the fracture toughness may increase if the pores cause the direction of crack propagation to deviate from the direction normal to the applied stress [82] In general, however, the fi'acture tbughness decreases as porosity increases due to the local applied stress acting over decreasing cross sectional area [85,82]. Fracture toughness decreases with increasing porosity (Section 3.4) for both sintered hydroxyapatite [54,56], and YBaQCu;,O7,x [78]. Microcracks induced by thermal expansion anisotropy or thermal expansion mismatch can decrease the fiacture toughness for many non-cubic ceramics [53]. For example microcracking due to thermal expansion anisotropy was deemed responsible for a 15% decrease in the fiacture toughness of B-silicon nitride as the average grain size increased [51]. Chou et al. [77] compared the elastic modulus and fi‘acture toughness of SiC platelet reinforced alumina for two different SiC platelet sizes. E and Klc for SiC platelet (24 um average diameter) reinforced alumina decreased 40% and 50% respectively as the volume fraction of SiC platelets increased from 0.0 to 0.4 [77]. Alumina reinforced by SiC platelets of 12 um average diameter exhibited a 65% increase in Klc for 0.4 volume fi'action SiC addition [77]. Thermal expansion mismatch resulted in microcracking between the large SiC platelets and the alumina matrix [77]. Microcracking due to thermal expansion mismatch and phase transformation resulted in toughening by crack shielding and a locally reduced elastic modulus for MoSiz/unstabilized ZrO2 composite [86]. K1c values for 20 volume percent ZrO2 /MoSi2 were over twice that of MoSi2 [86]. The absence of apparent grain size-property trends in the data from the sintered 119 hydroxyapatite specimens in this study implies that either 1) No microcracks exist or 2) Microcrack damage has saturated for the grain size range included in this study [36]. 3.6 Thermal Expansion Thermal expansion data for a high purity alumina reference specimen collected by identical methods as the sintered hydroxyapatite specimens prepared for this study, was compared to the American Institute of Physics Handbook [87] data for alumina (Table 3- 11). The average thermal expansion coefficients from 3 different ms with the alumina reference specimen were plotted against the handbook data (Figure 3-33). There was reasonable agreement up to ~ 4000 C, for thermal expansion coefiicient (or) taken on heating and cooling, with significant divergence between the heating and cooling curves above 6000 C (Figure 3-33). Table 3-1 1 Thermal expansion data for Alumina, from American Institute of Physics Handbook 2"” Edition [87]. Temperature (0 C) Parallel c-axis Perpendicular c-axis Polycrystalline (x1045/0C) (x10’6/0C) (x 104/OCT“ O 5.5 4.6 4.9 200 7.8 7.1 7.33 400 8.9 8.1 8.37 600 9.4 8.5 8.8 800 9.7 8.8 9.1 * calculated from lattice expansion data and equation 1-3. Thermal expansion coefficient data for sintered hydroxyapatite specimens numbers 120 .3 h J 1 "12 J— 2 . - e n I I I 810 "8 I'o I... - . mm ; -°.'x E 3 _ .0.0.I. x o _ 00 i- : I .9 e —— l' 3 .II to X ‘>% 4 .- DJ E .4. E 2 -— a: ': fir- l— 0 : l : l : 4T . § t l r + t + 4 l O 100 200 300 400 500 600 700 800 Temperature (C) I Heating 0 Cooling x Reference data [76] Figure 3-33 Thermal expansion coeficient for high purity alumina determined by dilatometry, as a fimction of temperature and compared with the American Institute of Physics Handbook Value [76] (Table 3-6). Dilatometry data taken from Appendix A. 121 4-10 and numbersl 1-12 on heating showed much more variability at temperatures greater than 600° C (Figure 3-34). The thermal expansion coefficients collected during cooling, for the same experimental runs, exhibit significant scatter at all temperatures and had a different shape than the heating data (Figure 3-35). The average thermal expansion coefficient, measured during heating, for all the hydroxyapatite specimens increases from 12.9 ppm at 1000 C to 19.1 ppm at 600° C and the average thermal expansion coefficient measured during cooling increases very little being 19.1 ppm at 100° C and 20.8 ppm at 600° C (Table 3-12). No correlation between the grain size, and the thermal expansion coefficients, measured at 400° C, are seen (Figure 3-36). The difi‘erence, Delta (A), in the thermal expansion coeflicient data, for the hydroxyapatite specimens, between the heating and cooling curves, for all the runs, is positive below 500° C and goes negative at temperatures greater than 500° C (Figure 3- 37). 122 N 01 J l N O l l .5 01 l l _s o l l Thermal Expansion Coefficient (ppm) 01 l L l r l A l .r l 1 l r v I l 0 100 200 300 400 500 600 700 800 Temperature #4 I#5 x#ss $1116 X#7 v #8 A #10 a #11 x #12 Q Ave. Figure 3-34 Thermal expansion coefficients, measured by dilatometry on heating, for sintered hydroxyapatite specimens numbers 4-8 and numbers 10-12. Data taken from Appendix A. 123 0.) O I I E -. e o #525 -r- . . . a: ~~ .0 21. E13 F, 8 XX :3 8‘ 83258] 220 -~ ‘B§§' EEA-e 23% §f§ .9 3:: WEE EVE .0 2 w .7 ¥§E i; 3m:XX:X :33; ‘8 to x6 x Lg- 15 .. « XxxngX '2 ’fix; _ E 'xZ XX x g 24 Zi' '03 -. o .C o *‘10 - t t . r 4. e : %4 9:55-1— 0 100 200 300 400 500 600 700 800 Temperature #4 I #5 EB #58 v #6 x #7 A #8 a #10 E #11 X #12 Figure 3-35 Thermal expansion coefficient data for sintered hydroxyapatite, from cooling dilatometry. The data taken from Appendix A. 124 Table 3-12 The averaged thermal expansion coefficients for sintered hydroxyapatite specimen numbers 4-8, and 10-12 by dilatometry. The delta is the difference between the cooling data and the heating data: Delta = a cool - a heat. Temperature (°C) Heating data (ppm) Cooling data (ppm) Delta (ppm) 100 12.89 19.07 6.18 200 13.75 18.41 4.66 300 15.12 17.78 2.66 400 16.34 17.42 1.08 500 17.52 18.78 1.26 - 600 19.79 20.77 0.98 700 20.20 17.52 -2.68 750 20.92 16.96 -3.96 125 16.5 ~— A I E Q. q.- 9; .3 16 a, I Q 55 I. O m- 0 I 8 3315.5“- C N a x ., LU E g 15 -+- - - "' .. I I 14.5 %1 i 4 : + .L : 4 + : : 1 2 3 4 5 6 7 8 Grain Size (Microns) Figure 3-36 Average thermal expansion coeflicient for sintered hydroxyapatite specimens #4-8 and #10-12 (heating to 600° C) as a function of the average grain size. The data was taken from Appendix A. 126 O 200 400 600 800 Temperature #4 .#5 E#5$'#6 x#7 A#8 8#1OE#11Z#12 Figure 3-37 Difierence in the thermal expansion coeflicient for sintered hydroxyapatite, Delta = on m- on ma , as a function of temperature. Dilatometry data taken from Appendix A. 127 3.7 Dielectric measurements 3.7.1 Circuit capacitance measurement and correction The circuit used to measure the relative permittivity was calibrated using the alumina reference specimen (ADS995, Coors) of known dielectric properties [72] (Table 3-13). The average measured capacitance (from three measurements at 45 kHz) of the circuit (Figure 2-7) with the alumina reference specimen inserted in the electrodes was 46.23 pF at 45 kHz. The capacitance of the alumina reference specimen in the electrodes (C ADS) was calculated from the known relative permittivity at 45 kHz and the geometry of the electrodes CADS = 28.89 pF (Equation 3-10) [88]. Cms=(€ode)/d Where 60 is the permittivity of free space, kd is the relative permittivity, A is the area of the (3-10) electrodes (corrected to add the width of the guard gap per ASTM D 150-95 [73]), and d is the specimen thickness. The circuit capacitance Cc was then established to be the difference in the measured capacitance Cm and the capacitance calculated for the reference specimen Cms- The circuit capacitance was determined to be Cc = CIn - CABS = 17.34 pF. The capacitance for the sintered hydroxyapatite specimens was determined by subtracting the circuit capacitance (17.34 pF) from the measured circuit capacitance. The circuit capacitance determined at 45 kHz was used at all frequencies. Table 3-13 Dielectric characteristics for Coors ADS 995 alumina substrate. Data from Tech Specs, Thin Film Substrates Technical Specifications 10-3-0897 from Coors Ceramics, Golden, CO [72]. Frequency Dielectric constant Loss Factor Q = 1/(Loss factor) 1 kHz 9.9 :1 0.003 333 1 MHz 9.9 :1 0.001 1000 128 3.7.2 Relative permittivity of hydroxyapatite The relative permittivity of the hydroxyapatite specimens was calculated from corrected capacitance measurements. The geometry of the electrodes and specimen thickness were used to calculate the relative permittivity of sintered hydroxyapatite by rearranging equation 3-3 [88] (Equation 3-11). kd=Cd/eoA (3-11) where the variables are as defined for equation 3-10. The fully dense hydroxyapatite relative permittivity (kw), was calculated for MGTl (Table 3-14) by a simple fit of the data to equation 1-9. For each specimen the difiemnce of the measured relative permittivity kd and the calculated km, fiom equation 1- 9 was calculated with the specimen’s relative density and the variable km. The differences were then summed. The fully dense hydroxyapatite relative permittivity (km) was incremented in equation 1-9 until the sum of the differences was at a minimum. The fiilly dense hydroxyapatite relative permittivity (kw) was also calculated for MGT2 and equation 1-10 (Table 3-14). The relative permittivity of sintered hydroxyapatite at 45 kHz decreased as the relative density decreased (Figure 3-3 8). As a function of the relative density the data for the hydroxyapatite specimens more closely matched MGT2 than MGTl (Figure 3-3 8). The relative permittivity of the sintered hydroxyapatite is stable over the measured frequency range (45 kHz - 7.3 MHz) (Figure 3-39, Figure 3-40). The relative permittivity of hydroxyapatite sintered with Mg2+ additives and partially decomposed, as determined by x-ray dim'action was measured by F anovich et al. [89]. The relative permittivity of the sintered hydroxyapatite specimens measured by 129 Table 3-14 Relative permittivity for sintered hydroxyapatite fiom this study and Fanovich et al. [89] and the fit parameters for the Maxwell-Garnet equations. \\ This study Fanovich et al. [89] ‘ Fit parameters Fit parameters R k MGTl MGT2 MGTI MGT2 K kccr 13.01 17.63 15.4 28.3 \ Difierence per 0.831 0.698 1.39 0.696 data point“ k Specimen Relative km Mg2+ Relative kTn Label Density weight % Density 44 0.952 11.63 1 0.897 13.62 14 0.951 13.55 0 0.856 12.61 43 0.947 11.62 3 0.846 9.81 45 0.937 12.05 5 0.830 9.45 16 0.936 12.64 \V‘ 42 0.936 12.47 50 0.931 11.53 49 0.924 12.02 46 0.918 I 1.63 41 0.769 7.94 36 0.658 5.62 33 0.579 4.93 x * The sum of the differences between the measured data and the fit equation divided by the number of total data points. 130 _s A l I _L N l T J O L l Relative Permittivity CD I 6- 4- 2—1 0.%i::r%*#*%1%r‘.:%‘rui 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 Relative Density I HAP — MGT1 — MGT2 Figure 3-38 Relative permittivity data fiom sintered hydroxyapatite and the fitted Maxwell-Garnet equations MGT1 (Equation 1-9) and MGT2 Equation (1- 10). The data was taken from Table 3-9. 131 14 a . 5 .. .3312 - 0 - I . x . E10 . - = x 2 Ba x E z: 8 BE 83 ¢_u i? 6 x I 9 ‘ l 1 i L J 1 l L I A l 1 l I ' I ' I ' l 1 r 1 l ' T 0 1000 2000 3000 4000 5000 6000 7000 Frequency (kHz) I ADSQQS O #14 x #16 V #33 A #36 E #41 E #42 Figure 3-39 Relative permittivity for the sintered hydroxyapatite specimen numbers 14, 16, 33, 41, 42 and the alumina reference material ADS995 as a function of fi'equency. The data was taken fi'om Appendix B. 132 13 g >. 312.5 I .=_: E 5 ‘ a, 12 v (L a) X €115 g .. i z E 11 A E l V 10.5 ¢ 1 I l I 1 I l I l I l I l I I 1 1 l I l v I I l 0 1000 2000 3000 4000 5000 6000 7000 Frequency (KHz) I#43. 1144211457 #46A #498 #50 Figure 3-40 Relative permittivity for sintered hydroxyapatite specimen numbers 43-46, 49 and 50 as a function of frequency. The data was taken from Appendix B. 133 Fanovich et al. [89] were greater than the sintered hydroxyapatite specimens in this study as a function of relative density. Fitting the data fiom Fanovich et al. [89] to MGT1 and MGT2 produced a closer match to MGT2 (Table 3-14). Sintered monoclinic hydroxyapatite of 97.6% relative density was determined to have a relative permittivity of 15.4 at 100 kHz by Ikoma et al. [26]. 134 4.0 Summary and Conclusions Commercially available hydroxyapatite powder can be sintered to 96% relative densities by simple uniaxial die pressing and pressure-less sintering at 13000 C. Final sintered density is a strong function of the pre-sintering densificatiort A linear relationship between final sintering density and uniaxial die pressure was determined. The final sintered density of hydroxyapatite specimens sintered at 1300° C for less than 120 minutes increased 0.27% in relative density for each additional MPa of pressure applied. An increase in the relative density of sintered hydroxyapatite to 98% required cold isostatic pressing before sintering at 13000 C. Average grain sizes for the hydroxyapatite specimens fell between 1.5 pm and 7.5 um and are a function of sintering time and temperature. There were no decomposition phases of hydroxyapatite detected by x-ray difli’action for hydroxyapatite specimens sintered at 1300° C with sintering times up to 11 hours and at 14000 C for as long as 4 hours. The phase stability of the hydroxyapatite is attributed to the slightly rich calcium content (36-3 7% by specific analysis) of the commercial hydroxyapatite powder and from the moisture present in the ambient atmosphere of the fiimaces used. The Vickers hardness values measured for the sintered hydroxyapatite specimens in this study were successfully fit, as a function of the volume fraction porosity (P), to the exponential decay equation (Hv = Ho e") with the zero porosity hardness value Ho = 6.00 i 0.7 GPa and decay constant b = 6.03. Fracture toughness of the sintered hydroxyapatite specimens in this study diverged fi'om the expected exponential decay function as the porosity increased. The measured 135 fi'acture toughness is 0.4-0.6 MPa m“2 for sintered hydroxyapatite specimens of 97% relative density. The fracture toughness of sintered hydroxyapatite increased to 2.5-2.7 MPa m“2 at 72-75% relative density. Published data fiom other researchers on the fracture toughness of sintered hydroxyapatite was successfully fit, in this study, to the exponential decay equation with nearly identical b values of 2.786 [54] and 2.812 [56]. Fracture toughness of specimens at 97% relative density were 1.34-1.5 MPa m”2 for hydroxyapatite whisker reinforced hydroxyapatite[56] and 1.01-1.2 MPa m"2 for dense unreinforced sintered hydroxyapatite [54,56,76]. Corrections to the indentation fi'acture toughness equation (Equation 2-6, Equation 3-3) for the decrease in the elastic modulus as the porosity increased decreased the calculated fracture toughness over the entire porosity range ( Kk(E(Pp)), Equation 3- 6). The exponential in equation 3-6 decreased K: to a greater degree as the porosity increased. ch increased as the porosity increased up to a 0.2 volume fiaction then decreased. KIc peaked at 121.6 MPa m"2 for specimens of 0.225 volume fi'action porosity. Comparing the fracture toughness measured by SENB to that measured by the indentation method for YBazCu3O7_x [78] suggested a further correction to the indentation fi'acture toughness equation K,C(E(Pp)). Assuming the empirical constant is a function of porosity C(Pp), K1c measured by the indentation method can be equated to Klc measured by SENB for the YBaQCu3O-,_x specimens. C(Pp) added to K,c(E(Pp)) produced K,c(C(Pp),E(Pp)) (Equation 3-9). Applying ch(C(Pp),E(Pp)) to the indentation of the sintered hydroxyapatite specimens prepared for this study increased Klc to 0.6-1.0 MPa m"2 at low porosity (~ 0.03 volume fiaction porosity). Klc peaked at 1.8-2.0 MPa m"2 for volume fraction porosities 0.06 to 0.17 before decreasing. 136 Both corrections to the indentation fiacture toughness equation (Equation 3-6, and 3-9) when applied to the sintered hydroxyapatite indentation data produced increasing Klc at low and moderate porosities (0.02-0.15) before falling off. A small percentage (5 out of 2 1) of the Kk/Porosity studies found in the literature report other than decreasing K,c as porosity increases. The hysteresis seen in the heating and cooling curves of sintered hydroxyapatite from the dilatometer was determined to be an artifact of the apparatus and not due to microcracking. The same hysteresis was seen for dilatometry data for the non- microcracked alumina specimen. The thermal expansion coefficients for sintered hydroxyapatite, averaged for the heating data for all the hydroxyapatite specimens tested, equals 12.89 x10‘6/°C at 100° C , 13.75 x10‘/°C at 2000 C, 16.34 at 400° C and 19.79 x 1015/ °C at 600° C. The dilatometer was not accurate above 6000 C based on the comparison between the handbook data [71] and the dilatometer values for the high purity alumina specimen above 6000 C. The maximum relative dielectric permittivity of sintered hydroxyapatite in this study was k = 13.55 for a specimen of 95.1% relative density. The minimum relative dielectric permittivity k = 4.93 for a specimen at 58.0% relative density. Maxwell-Garnet mixing laws were used to determine the zero porosity relative permittivity and the functional relationship between the relative permittivity and porosity of sintered hydroxyapatite specimens. The first mixing law (MGT 1) assumes spherical pores in a ceramic matrix and was the least successful in fitting the data fiom this study. MGT2 in which spherical ceramic particles are in a matrix of air fit the data more successfully. MGT1 predicted a zero porosity relative permittivity of 13.01. MGT2 predicted k0 = 137 17.63 fiom the data in this study. 4.1 Further Research The indentation technique for fracture toughness measurements used in this study, though used successfully for sintered hydroxyapatite by Suchanek et al. [56] needs fiu'ther scrutiny. The size of the indentations grew as the sintered hydroxyapatite specimens became more porous, as expected: The larger indents, however, failed to produce the increased extent of the half-penny cracks. Possrble mechanisms that would limit the indentation cracks in porous hydroxyapatite are densification below the indent and crack bridging between pores. Densification below the indent tip under load would absorb energy and disrupt the plastic/elastic interface responsible for indentation crack grth [90]. Crack bridging results in greater surface areas formed from halfpenny cracks of seemingly the same length. 138 APPENDIX A Appendix A Thermal expansion data. The thermal expansion data collected at 25 °C intervals fiom 50 °C to 800 0C for both the heating and cooling curves. The data was taken for sintered hydroxyapatite samples numbers (4-8, 10 and 11) along with two sintered alumina specimens. Table A1 Thermal expansion data for Hydroxyapatite samples number 4, 5 with no dwell at 800 °C, and 5 with a pause at 800 °C. The data is used in Figures 3-26 to 3-27. Temperature 4 (um/m oC ) 5 (um/m 0C ) 5 soak (um/m oC) (0C) Heat Cool Heat Cool Heat Cool 50 12.2 0 11.4 N/A 12.0 N/A 75 12.7 3 12.7 N/A 13.1 15.8 100 12.6 0 13.0 N/A 13.5 15.4 125 12.6 , 13.7 13.0 N/A 13.9 14.8 150 13.1 16.1 13.4 20.3 14.4 15.1 175 13.4 16.8 13.9 21.9 15.1 16.2 200 14 N/A 14.2 22.4 15.3 15.3 225 13.8 17.8 14.8 23.4 15.6 16.2 250 14.1 18.8 14.6 25.0 14.4 16.6 275 14.5 17.2 15.5 26.5 15.0 15.9 300 14.2 17.7 16.5 18.0 15.9 16.6 325 14.9 17.0 16.3 17.3 15.5 15.5 350 15.0 18.6 16.5 N/A 15.9 15.7 375 15.4 18.5 16.9 21.6 15.9 15.1 400 16.6 19.3 17.5 16.1 17.8 16.5 425 16.1 19.8 17.1 17.1 17.5 17.4 450 16.2 20.7 17.6 23.9 15.8 18.4 475 16.1 20.2 17.2 21.3 17.2 19.1 139 APPENDIX A Table A1 (cont’d) Temperature 4 (um/m °C ) 5 (um/m °C ) 5 soak (um/m °C ) (0C) Heat Cool Heat Cool Heat Cool 500 17.0 19.6 18.0 23.8 17.7 19.6 525 17.4 20.2 18.3 20.8 19.2 19.8 550 16.6 21.3 19.2 21.0 19.0 19.8 575 18.1 18.9 20.4 19.0 18.8 20.7 600 18.0 20.4 20.5 18.6 15.9 21.8 625 20.0 20.9 22.6 18.5 21.7 21.5 650 17.0 18.1 21.1 17.6 21.4 19.9 675 18.8 18.2 21.7 15.9 22.1 18.4 700 19.0 15.5 23.2 15.5 20.0 19.7 725 18.2 14.5 24.0 20.9 20.9 19.2 750 17.9 16.3 21.1 10.9 22.1 18.7 775 19.5 15.8 19.7 11.8 22.9 17.2 800 20.7 N/A 20.0 9.64 N/A N/A Table A2 Thermal expansion data for Hydroxyapatite samples number 7, 8, and 10 fiom Figures 3-26 to 3-27. Temperature 6 (um/m °C ) 7 (um/m °C ) 8 (um/m oC) (0C) Heat Cool Heat Cool Heat Cool 50 12.7 13.7 11.6 75 12.8 13.0 12.5 100 13.0 12.9 12.5 125 12.8 16.9 13.0 15.7 13.2 17.2 150 13.3 17.4 13.6 15.5 13.4 18.2 175 13.9 18.2 13.7 15.8 9.58 18.6 200 13.5 15.6 14.1 16.0 10.4 16.7 140 Appendix A Table A2 continued Temperature 6 (um/m °C ) 7 (um/m °C ) 8 (um/m °C ) (0C) Heat Cool Heat Cool Heat Cool 225 14.5 17.9 13.9 16.1 11.6 17.7 250 14.7 18.9 14.3 16.3 13.2 17.5 275 15.4 18.6 14.9 16.3 13.7 13.3 300 14.9 18.5 14.9 16.7 15.3 19.1 325 15.7 17.2 15.1 16.4 15.7 18.5 350 16.6 17.4 14.8 16.7 15.4 N/A 375 17.3 19.2 14.9 16.8 15.9 19.5 400 14.9 15.3 15.4 17.0 15.6 20.3 425 16.5 18.9 14.8 14.2 17.0 19.6 450 17.5 . 18.8 15.6 13.3 16.7 21.1 475 18.4 13.5 15.0 18.4 17.5 21.6 500 18.3 16.2 15.8 17.9 16.9 19.6 525 19.9 21.0 15.7 17.7 18.3 19.7 550 19.6 20.0 17.1 19.5 16.8 19.4 575 17.8 20.6 18.2 19.8 18.1 20.5 600 20.9 20.6 18.9 21.2 19.0 19.8 625 24.0 21.4 23.4 23.4 27.8 17.5 650 21.3 20.0 20.2 19.4 16.8 18.5 675 21.1 18.5 20.1 18.5 23.51 18.3 700 20.0 17.7 22.3 15.6 18.9 18.4 725 19.7 16.4 22.1 15.2 19.6 16.4 750 20.2 15.9 20.5 14.6 18.9 N/A 775 21.6 15.2 20.9 15.4 19.9 18.5 800 27.2 N/A 21.1 14.0 28.9 N/A 141 APPENDIX A Table A3 Thernml expansion data for Hydroxyapatite samples numbers 10, 11, and 12 from Figures 3-26 to 3-27. Temperature 10 (um/m oC ) 11 (um/m °C) 12 (um/m °C) (°C) Heat Cool Heat Cool Heat Cool 50 12.6 N/A 12.8 N/A 11.9 18.8 75 12.8 N/A 13.2 18.3 12.3 23.1 100 13.0 N/A 13.4 19.1 12.1 22.7 125 13.8 19.1 13.4 19.1 11.4 21.2 150 14.4 18.2 13.9 14.3 11.0 21.1 175 14.4 19.4 14.2 17.6 12.0 N/A 200 14.7 20.5 13.2 20.8 13.5 20.0 225 14.4 19.0 13.7 17.3 14.0 19.7 250 15.1 13.3 13.9 16.2 14.4 18.0 275 15.8 18.8 14.1 17.3 14.3 18.6 300 15.6 18.4 15.0 17.4 14.6 17.6 325 15.9 19.8 15.5 18.7 14.9 N/A 350 17.1 21.7 16.1 17.0 14.9 17.1 375 16.9 18.4 16.3 19.1 15.3 17.5 400 17.8 17.8 16.7 16.8 15.4 17.7 425 17.5 18.1 16.2 17.9 16.2 13.3 450 18.7 13.4 17.2 18.2 16.5 14.7 475 16.3 19.1 17.9 19.5 16.8 16.2 500 18.9 N/A 18.0 18.6 16.5 14.9 525 19.6 20.3 18.2 18.7 16.9 16.1 550 19.7 21.3 18.9 19.7 16.7 17.1 575 20.3 20.5 19.9 20.3 18.4 17.7 600 22.1 22.4 18.9 20.9 17.5 21.2 142 APPENDIX A Table A3 continued 625 25.0 23.8 22.9 19.8 20.5 19.6 650 21.5 20.7 20.6 19.4 17.6 17.5 675 21.1 22.5 21.5 22.3 17.3 16.7 700 20.7 20.8 19.8 18.3 16.6 16.2 725 20.8 21.1 19.8 21.0 19.5 10.3 750 20.6 21.5 20.1 21.5 17.9 16.3 775 21.2 19.8 22.3 N/A 18.1 10.1 800 26.9 18.5 23.2 N/A 19.6 N/A Table A4 Thermal expansion data for the alumina sample number 1 with 1 g weight, alumina number 1 with 5 g weight (denoted 1a) and alumina sample number 2. Temperature 1 (um/m _°C ) 1a (um/m oC ) 2 (um/m oC ) ( C) Heat Cool Heat Cool Heat Cool 50 5.24 4.92 5.27 75 5.66 5.40 5.62 100 5.80 5.84 5.76 5.49 125 6.26 8.77 6.06 6.09 150 6.57 8.38 6.38 6.41 175 6.73 8.78 6.74 6.85 200 7.09 8.24 7.03 7.03 6.76 8.90 225 7.14 8.53 7.38 7.36 250 7.42 8.71 7.28 7.58 275 7.56 8.92 7.77 7.69 300 7.8 8.77 8.12 7.95 8.61 9.20 325 7.86 8.77 8.31 8.21 350 8.35 8.96 8.90 8.39 143 APPENDIX A Table A4 continued Temperature #1 (1.1an °C ) #1a(um/m °C ) #2 (um/m °C ) (0C) Heat Cool Heat Cool Heat Cool 375 9.36 9.04 8.72 8.73 400 8.9 9.20 9.00 9.09 9.15 9.36 425 9.15 9.57 9.15 9.18 450 9.4 9.25 9.36 9.1 475 9.95 9.58 9.40 8.97 500 9.93 8.95 9.82 9.89 11.0 8.81 525 9.97 9.35 9.95 9.69 550 11.0 8.99 10.4 10.1 575 10.5 9.54 9.66 10.8 600 10.5 q 8.67 9.96 10.3 9.16 9.40 625 10.5 8.62 11.0 11.5 650 10.3 8.32 10.3 10.9 675 11.3 8.48 10.9 10.7 700 10.7 7.31 11.5 11.5 12.2 8.83 725 9.24 7.36 9.99 11.5 750 11.8 5.96 12.3 11.5 775 10.2 7.26 11.1 11.5 144 APPENDEX A Table A5 The averaged therrml expansion data for sintered hydroxyapatite (from Tables A1,A2 and A3) and alumina (fiom Table A4). Temperature HAP (um/m °C ) Alumina (um/m °C ) (0C) Heat Cool Heat Cool 50 12.32 18.8 5.14 5.27 75 12.79 19.07 5.56 5.62 100 12.89 19.07 5.72 5.76 125 13.01 17.21 6.14 7.43 150 13.39 17.36 6.45 7.40 175 13.35 18.06 6.77 7.82 200 13.66 18.41 6.98 8.06 225 14.03 18.34 7.29 7.95 250 14.30 17.84 7.43 8.15 275 14.80 18.06 7.67 8.31 300 15.21 17.78 8.12 8.64 325 15.50 17.55 8.13 8.49 350 15.81 17.74 8.55 8.68 375 16.09 18.41 8.94 8.89 400 16.41 17.42 9.04 9.22 425 16.54 17.37 9.16 9.38 450 16.87 18.06 9.29 9.18 475 16.93 18.77 9.44 9.28 500 17.46 18.78 10.16 9.22 525 18.17 19.37 9.87 9.52 550 18.18 19.90 10.50 9.55 575 18.89 19.78 10.32 10.17 600 19.08 20.77 9.98 9.46 145 Table A5 continued APPENDEX A Temperature HAP (um/m oC ) Alumina (urn/m °C ) (0C) Heat Cool Heat Cool 650 19.72 19.01 10.5 9.61 675 20.80 18.81 10.97 9.59 700 20.06 17.52 1 1.48 9.21 725 20.51 17.22 10.24 9.43 750 19.92 16.96 1 1.87 8.73 775 20.68 15.48 10.93 9.38 800 23.45 14.05 N/A N/A 146 APPENDIX B Appendix B The dielectric data The calculated relative permittivity and the measured quality factor Q, for the Alumina substrate (ADS 995) and hydroxyapatite specimens numbers 14, 16, 33, 36, 41- 46, 49, 50. The data was taken in the fi'equency range 45 kHz to 7.3 MHz. Table B1 Dielectric data for alumina substrate (1.016 mm in thickness). Frequency (kHz) Relative permittivity Q 45.0 9.9 333 72.17 9.3 244 101.62 8.8 268 160.77 10.2 235 307.17 8.7 196 702.20 9.9 309 1641.90 9.9 217 31 12.91 9.6 246 7302.00 1 1.4 195 147 APPENDIX B Table B1 Dielectric data for hydroxyapatite specimen number 14 (2.845 mm in thickness). Frequency (kHz) Relative Permittivity Q 45.0 13.6 52.6 72.17 13.3 25.4 101.62 13.2 30.3 160.77 14.1 35.9 307.17 12.2 42.5 702.20 13.1 50.7 1641.90 12.1 55.9 3112.91 10.4 58.1 7302.00 1 1.9 69.2 Table B1 Dielectric data for hydroxyapatite specimen number 16 (2.969 mm in thickness). Frequency (kHz) Relative Permittivity Q 45.0 12.6 56.2 72.17 12.6 27.0 101.62 10.6 35.8 160.77 12.9 37.1 307.17 12.8 44.4 702.20 13.7 64.4 1641.90 10.7 84.8 3112.91 8.9 80.8 7302.00 9.9 1 15 148 APPENDIX B Table B1 Dielectric data for hydroxyapatite specimen number 33 (1.123 mm in thickness). Frequency (kHz) Relative Permittivity Q 45.0 4.9 185 72.17 4.7 105 101.62 4.1 96.8 160.77 4.7 82.8 307.17 4.3 108 702.20 5.4 93.7 1641.90 5.4 84.4 31 12.91 4.6 104 7302.00 4.5 132 Table B1 Dielectric data for hydroxyapatite specimen number 36 (1.730 mm in thickness). Frequency (kHz) Relative Permittivity Q 45.0 5.6 186 72.17 3.9 87.7 101.62 4.8 93.9 160.77 5.5 77.6 307.17 3.8 90.8 702.20 5.2 79.2 1641.90 5.4 65.6 3112.91 5.2 84.6 7302.00 4.6 88.2 149 APPENDIX B Table B1 Dielectric data for hydroxyapatite specimen number 41 (0.866 mm in thickness). Frequency (kHz) Relative Permittivity Q 45.0 7.9 156 72.17 7.8 108 101.62 7.5 l 14 160.77 7.8 95.4 307.17 7.4 101 702.20 7.8 86.3 1641.90 8.3 71.7 3112.91 8.1 61.4 7302.00 8.2 42.4 Table B1 Dielectric data for hydroxyapatite specimen number 42 (1.113 mm in thickness). Frequency (kHz) Relative Permittivity Q 45.0 12.5 37.2 72.17 12.3 26.3 101.62 11.5 28.7 160.77 11.9 33.1 307.17 11.5 41.8 702.20 1 1.7 49.9 1641.90 12.3 56.3 3112.91 11.6 72.3 7302.00 12.6 72.6 150 APPENDIX B Table B1 Dielectric data for hydroxyapatite specimen number 43 (1.336 mm in thickness). Frequency (kHz) Relative Permittivity Q 45.0 11.6 127 72.17 10.6 70.4 101.62 11.0 92.7 160.77 1 1.0 125 307.17 10.7 87.8 702.20 1 1.0 134 1641.90 10.9 139 3112.91 11.2 133 7302.00 13.0 119 Table B1 Dielectric data for hydroxyapatite specimen number 44 (1.214 mm in thickness). Frequency (kHz) Relative Permittivity Q 45.0 11.6 110 72.17 11.6 78.6 101.62 11.7 87.2 160.77 11.0 95.8 307.17 10.9 134 702.20 1 1.0 155 1641.90 12.1 165 3112.91 12.2 177 7302.00 12.9 148 151 APPENDIX B Table B1 Dielectric data for hydroxyapatite specimen number 45 (1.610 mm in thickness). Frequency (kHz) Relative Permittivity Q 45.0 12.1 229 72.17 11.3 148 101.62 ’ 10.6 168 160.77 11.6 135 307.17 11.2 152 702.20 11.2 173 1641.90 12.2 164 3112.91 11.7 156 7302.00 12.5 131 Table B1 Dielectric data for hydroxyapatite specimen number 46 (1.651 mm in thickness). Frequency (kHz) Relative Permittivity Q 45.0 11.6 184 72.17 1 1.3 97.3 101.62 11.6 106 160.77 11.4 158 307.17 1 1.1 108 702.20 10.4 179 1641.90 1 1.1 1 12 3112.91 10.9 138 7302.00 11.9 11 1 152 APPENDIX B Table B1 Dielectric data for hydroxyapatite specimen number 49 (1.580 mm in thickness). Frequency (kHz) Relative Permittivity Q 45.0 12.0 245 72.17 1 1.1 154 101.62 1 1.4 146 160.77 1 1.3 158 307.17 1 1.2 157 702.20 10.6 199 1641.90 10.8 166 3112.91 11.1 127 7302.00 11.0 157 Table B1 Dielectric data for hydroxyapatite specimen number 50 (1.595 mm in thickness). Frequency (kHz) Relative Permittivity Q 45.0 11.5 188 72. 17 l 1.2 136 101.62 1 1.2 140 160.77 10.1 93.1 307.17 1 1.0 103 702.20 10.8 132 1641.90 1 1.0 123 3112.91 11.2 106 7302.00 11.2 150 153 [11 [2] [3] [4] [5] [6] [7] [31 [9] [101 References Larry L. Hench, “Bioceramics”. Journal of the American Ceramic Society, 81 [7], pp.1705-1727 (1998). A.M.J.H. Sueter, “Existence Region of Calcium Hydroxylapatite and the Equilhbrium with Coexisting Phases at Elevated Temperatures”, Reactivity of Solids, pp 806-812, J.S. Anderson, M.W. Roberts, Eds., Chapman and Hall, London (1972). P. V. Riboud, “Composition and stability of apatites in the system CaO-P205 - Iron oxide - H20 at high temperatures”, Annales de chimie, 8 pp 381-390 (1973). J .C. Trombe, G. Monte], “Some Features of the Incorporation of Oxygen in Different Oxidation States in the Apatitic Lattice-I”, J. Inorg. Nucl. Chem. Vol. 40, pp 15-21, (1978). R. M. H. Verbeek, H. J. M. Heiligers, F. C. M. Driessens, and H. G. Schaeken, ‘Efi‘ect of Dehydration of Calcium Hydroxylapatite on its Cell Parameters”, Z. anorg. allg. Chem. 466, pp 76-80, (1980). Jiming Zhou, Xingdong Zhang, Jiyong Chen, “High temperature characteristics of synthetic hydroxyapatite”, Journal Of Materials Science: Materials in Medicine, 4 pp 83-85 (1993). A. Royer, J .C. Viguie, M. Heughebaert, J.C. Heughebaert: “Stoichiometry of Hydroxyapatite: influence on the Flexural Strength”, Journal of Materials Science: Materials in Medicine, 4, pp76-82, (1993). Kimihiro Yamashita, Kazuhisa Kitagaki, and Takao Umegaki: “Thermal Instability and Proton Conductivity of Ceramic Hydroxyapatite at High Temperatures”, Journal of the American Ceramic Society, 78 [5], pp.1 191 - 1197 (1995). Pauchiu B. Wang, T.K. Chaki: “Sintering Behaviour and Mechanical Properties of Hydroxyapatite and Dicalcium Phosphate”, Journal of Materials Science: Materials in Medicine 4 pp.150-158, (1993). A.J. Ruys, C.C. Sorrel, A. Brandwood, B.K. Milthorpe: “The Effects of Sintering Atmosphere on the Chemical Compatibility of Hydroxyapatite and Particle Additives at 1200° C”, Journal of Materials Science: Materials in Medicine, 6, pp297-301, (1995). 154 [11] [12] [13] [141 [15] [16] [17] [131 [19] [20] [21] Anna Slosarczyk, Ewa Stobierska, Zofia Paskiewicz, Mark Gawlicki: “Calcium Phosphate Materials Prepared from Precipitates with Various Calcium:Phosphorus Molar Ratios”, Journal of the American Ceramic Society, 79 [10], 2539-2544 (1996). Wojciech L. Suchanek and Masahiro Yoshimura, “Preparation of Fibrous, Porous Hydroxyapatite Ceramics from Hydroxyapatite Whiskers”, Journal of the American Ceramic Society, 81, [3], pp 765-767, (1998). Kazuo Kondo, Masahiko Okuyama, Hidetoshi Ogawa and Yoshimasa Shibata. “Preparation of High-Strength Apatite Ceramics”, Communications of the American Ceramic Society, November 1984, pp C222-223. W. Fix, H. Heymann and R Heinke, “Subsolidus Relations in the System 2CaO SiOz-3CaO P20, “, Journal of the American Ceramic Society, 52, [6] pp 46-47 (1969). J .C. Elliott, Structure and Chemistry of the Apatites and Other Calcium Orthophosphates, Elsevier, (1994). Sir Lawrence Bragg, G.F. Claringbull, W.H. Taylor: Crystal Structures of Minerals, Cornell University Press. Ralph W.G. Wyckofl‘: Crystal Structures Volume 3 Inorganic Compounds RIM X,),,I§K (M,1 Xp)y, Hydrates and Ammoniates, Second Edition, Interscience Publishers. D.E.C. Corbridge; Studies in Inorganic Chemistry 20, Phosphourous An Outline of its Chemistry and Uses, Fifth Edition, Elsevier (1995). Tsuyoshi Kijima and Masayuki Tsutsumi, “Preparation and Thermal Properties of Dense Polycrystalline Oxyhydroxyapatite”, Journal of the American Ceramic Society, 62 [9-10], pp.455-460 (1979) M. Akao, H. Aoki, K. Kato, “Mechanical Properties of Sintered Hydroxyapatite for Prosthetic Applications”, Journal of Material Science, 16, pp.809-812 (1981). Keizo Uematsu, Masaki Takagi, Tsunehiro Honda, Nozumu Uchida and Katsuichi Saito: “Transparent Hydroxyapatite Prepared by Hot Isostatic Pressing of Filter Cake”, Journal of the American Ceramic Society, 72 [8], pp.1476-1478, (1989). 155 [22] [23] [241 [251 [26] [27] [28] [29] [30] [31] [32] A. Cuneyt Tas, F. Korkusuz, M. Timucin, N. Akkas: “An Investigation of the Chemical Synthesis and High Temperature Sintering Behaviour of Calcium Hydroxyapatite (HA) and Tricalcium Phosphate (TCP) Bioceramics”, Journal of Materials Science: Materials in Medicine 8 pp.91- 96, (1997). Yi Fang, Dinesh K. Agrawal, Della M. Roy, and Rustum Roy: “Fabrication of Porous Hydroxyapatite Ceramics by Microwave Processing”, Journal of Material Research, Vol. 7 No. 2, pp.490-494 (1992). J .C. Elliott, P.E. Mackie, RA. Young: “Monoclinic Hydroxyapatite”, Science, Vol 180, pp.1055-1057 (1973). H.B. Van Rees, M. Mengeot, E. Kostiner: ‘Monoclinic-Hexagonal Transition in Hydroxyapatite and Deuterhydroxyapatite Single Crystals”, Mat. Res. Bull. Vol 8, No. 11, pp.1307-1309 (1973). Toshiyuki Ikoma, Atsushi Yarmzaki: “Preparation and Dielectric Property of Sintered Monoclinic Hydroxyapatite”, Journal of Materials Science Letters 18, pp.1225-1228 (1999). Hiroyuki Suda, Masatomo Yashirna, Masoto Kakihana, and Masahiro Yoshimura, “Monoclinic H Hexagonal Phase Transition in Hydroxyapatite Studied by X-ray Powder Diffraction and Differential Scanning Calorimetry Techniques”, J. Phys. Chem, 99, pp 6752-6754, (1995). G. R. Fischer, P. Bardhan, J. E. Geiger, “The lattice thermal expansion of hydroxyapatite”, Journal of Materials Science Letters, 2, pp 577-578, (1983). W.G. Perdok, J Chfistofiersen and J. Arends, “The Thermal Lattice Expansion of Calcium Hydroxyapatite”, Journal of Crystal Growth, 80, pp 149-154, (1987). M. Jarcho, C.H. Bolen, M.B. Thomas, J. Bobick, J.F. Kay, R.H. Doremus, “Hydroxylapatite synthesis and characterization in dense polycrystalline form”, Journal of Materials Science, 11, pp 2027-2035, (1976). Je-Won Choi, Young-Min Kong and Hyoun-Ee Kim, In-Seop Lee, “Reinforcement of Hydroxyapatite Bioceramic by Addition of Ni3A1 and A1203“, Journal of the American Ceramic Society, 81, [7], pp 1743-1748, (1998). M. Mathew, L.W. Schroeder, B. Dickens and WE. Brown: “The Crystal Structure of o1-Ca3(PO4 2”, Acta Cryst. B33, ppl325-1333 (1977). 156 [33] [34] [35] [36] [37] [33] [39] [40] [41] [42] [43] [44] B. Dickens, L.W. Schroeder, and WE. Brown, “Crystallographic Studies of the Role of Mg as a Stabilizing Impurity in [i-Ca3 (P04)2 I. The Crystal Structure of Pure B-Ca3 (PO,)2", Journal of Solid State Chemsitry 10, pp 232-248, (1974). B. Dickens, W.E. Brown, G.J. Kruger and J .M.. Stewart, “Ca4(PO4)zO, Tetracalcium Diphosphate Monoxide. Crystal Structures and relationships to Ca, (PO4)3OH and K3Na(SO4)2", Acta Cryst. B29, pp 2046-2056, (1973) W.D. Kingery, H.K. Bowen, D.R. Uhlrnann, Introduction to Ceramics, Second Edition, John Wiley & Sons, 1976. J .1 Cleveland and RC. Bradt, “Grain Size/Microcracking Relations for Pseudobrookite Oxides”, Journal of the American Ceramic Society, 61, [1 1-12], pp 478-481, (1978). Bernard J. Hockey, “Plastic Deformation of Aluminum Oxide by Indentation and Abrasion”, Journal of the American Ceramic Society, 54, [5], pp 223-231, (1971). Brian Lawn, Rodney Wilshaw, “Review Indentation Fracture: principles and applications”, Journal of Materials Science, 10, pp 1049-1081, (1975). AG. Evans and EA. Charles, “Fracture Toughness Determinations by Indentation”, Journal of the American Ceramic Society, 59, [7-8], pp 371-372, (1976) G.R. Anstis, P. Chantikul, B.R. Lawn, and DB. Marshall, “A Critical Evaluation of Indentation Techniques for measuring Fracture Toughness: 1, Direct Crack Measurements”, Journal of the American Ceramic Society, 64, [9], pp 533-538, (1981) James Lankford, “Indentation microfiacture in the Palmqvist crack regime: implications for fiacture toughness evaluation by the indentation method”, Journal of Materials Science Letters, 1, pp 493-495, (1982). M.T. Laugier, “Palmqvist cracking in WC-Co composites”, Journal of Material Science Letters, 4, pp 207-210, (1985). Roy W. Rice, Carl Cm. Wu, and Fred Borchelt, “Hardness-Grain-Size Relations in Ceramics”, Journal of the American Ceramic Society, 77, [10], pp 2539-2553, (1994). Andreas Krell and Paul Blank, “Grain Size Dependence of Hardness in Dense Submicrometer Alumina”, Journal of the American Ceramic Society, 78, [4], (1995). 157 [45] [461 [47] [43] [49] [50] [51] [52] [53] [54] [55] M. Desmaison-Brut and J. Montintin, F. Valin and M. Boncoeur, “Influence of Processing Conditions on the Microstructure and Mechanical Properties of Sintered Yttrium Oxides”, Journal of the American Ceramic Society, 78, [3], pp 716-722, (1995). M.J. Mayo, R.W. Siegel and Y.X. Liao, W.D. Nix, “Nanoindentation of nanocrystalline ZnO”, Journal of Materials Research, 7, (4) pp 973-979, (1992). M.J. Mayo, KW. Siegel and A. Narayanasamy, W.D. Nix, “Mechanical properties of nanophase TiO2 as determined by nanoindentation”, Journal of Materials Research, 5, (5), pp 1073-1082, (1990). J. Ricote, L. Pardo, B. Jimenez, “Mechanical characterization of calcium-modified lead titanate ceramics by indentation methods”, Journal of Materials Science pp 3248-3254, (1994). Dong Soo Baik, Kwang $00 No and John Soung-Soon Chun, “Mechanical Properties of Mica Glass-Ceramics”, Journal of the American Ceramic Society, 78, [5], pp 1217-1222, (1995). Jie Luo, R. Stevens, Wai Lo, A.M. Campbell, “Anisotropic plastic and elastic- deformation in highly textured superconducting (Bi, Pb)2 Sr2 Ca2 Cu3 Ox ceramics”, Journal of Materials Science, 30, (12), pp 3050-3056, (1995). Naoto Hirosaki and Yoshio Akimune, “Effect of Grain Grth of B-Silicon Nitride on Strength, Weibull Modulus, and Fracture Toughness”, Journal of the American Ceramic Society, 76, [7], pp 1892-1894, (1993). KW. Rice, R.C. Pohanka, and W.J. McDonough, “Effect of Stresses fiom Thermal Expansion Anisotropy, Phase Transformations, and Second Phases on the Strength of Ceramics”, Journal of the American Ceramic Society, 63, [1 1-12], pp 703-710, (1980). R. W. Rice, “Grain size and porosity dependence of ceramic fiacture energy and toughness at 22° C”, Journal of Materials Science, 31, pp 1969-1983, (1996). G. DeWith, H.J.A. Van Dijk, N. Hattu, K. Prijs, “Preparation, microstructure and mechanical properties of dense polycrystalline hydroxy apatite”, Journal of Material Science, 16, pp 1592-1598, (1981). R. Halouni, D. Bernache—Assolant, E. Champion, A. Ababou, “Microstructure and related properties of hot pressed hydroxyapatite ceramics”, Journal of Material Science: Materials in Medicine, 5, pp 563-568, (1994). 158 [56] [57] [58] [59] [60] [61] [621 [63] [64] [65] [661 [67] Wojciech Suchanek, Masatomo Yashirna, Masato Kakihana and Masahiro Yoshimura, “Hydroxyapatite/Hydroxyapatite-Whisker Composites without Sintering Additives: Mechanical Properties and Microstructural Evolution”, Journal of the American Ceramic Society, 80, [11], pp 2805-2813, (1997). R.W. Rice, “Evaluation and extension of physical property-porosity models based on minimum solid area”, Journal of Materials Science, 31, pp 102-118, (1996). R. W. Rice, “Comparison of physical property-porosity behaviour with minimum solid area models”, Journal of Material Science, 31, pp 1509-1528, (1996). TH. Nielson and M.H. Leipold, “Thermal Expansion in Air of Ceramic Oxides to 22000 C”, Journal of the American Ceramic Society, 46, [8], pp 381-387, (1963). KL. Coble and W.D. Kingery, “Effect of Porosity on Physical Properties of Sintered Alumina”, Journal of the American Ceramic Society, 39, [11], pp 377- 3 85, ( l 956). W.R. Manning, 0. Hunter Jr., F .W. Calderwood, and D.W. Stacy, “Thermal Expansion of szOs", Journal of the American Ceramic Society, 55, [7], pp 342- 347, (1972). J .A. Kuszyk and RC. Bradt, “Influence of Grain Size on Effects of Thermal Expansion Anisotropy in MgTiZOS", Journal of the American Ceramic Society, 56, [8], pp 420-423, ( 1973). Fred J. Parker and Roy W. Rice, “Correlation between Grain Size and Thermal Expansion for Aluminum Titanate Materials”, Communications of the American Ceramic Society, 72, [12], pp 2364-2366, (1989). ED. Case, J.R Smyth, O. Hunter Jr., “Microcrack Healing During Temperature Cycling of Single Phase Ceramics”, Fracture Mechanics of Ceramics, Volume 5, Plenum Press (1983). JP Calarne, A. Birman, Y. Carmel, D. Gershon, and B. Levush, A.A. Sorokin and V.E. Semenov, D. Dadon, L.P. Martin and M. Rosen, “A dielectric mixing law for porous ceramics based on fi'actal boundaries”, Journal of Applied Physics, 80, (7), pp 3992-4000, (1996). Particle size distribution test (16 March 2000), Hydroxyapatite powder item number C-2071-1, Batch number X16907-T, run number 13112, measurement number 364, Batch number X21949, run number 13113, measurement number 372, Cerac Incorporated, Milwaukee, WI. Type R N.I.S.T Monograph 175 Revised to ITS-90, “Revised Thermocouple Reference Tables”, Published by Omega Engineering, Stamford CT. 159 [68] [69] [70] [711 [721 [731 [74] [751 [76] [771 [78] [791 [801 Emscope SC500 instruction booklet. Powder diffraction file for inorganic phases, American Society for Testing and Materials, Philadelphia, PA, 1967. B.D. Cullity, Elements of X-Ray Diffraction Second Edition, Addison-Wesley Publishing Company, Inc. 1978. J .B. Wachtrnan, Mechanical Properties of Ceramics, pp 82-84, John Wiley and Sons, New York, 1996. Thin Film Substrates Technical Specifications 10-3-0897, Coors Ceramics Company, Golden CO. 1998 ASTM Designation D 150 - 95, “Standard Test Methods for AC loss Characteristics and Permittivity (Dielectric Constant) of Solid Electrical Insulation”, pp26-44. Hui Lu, Zhe Qu, Yanchun Zhou, “Preparation and mechanical properties of dense polycrystalline hydroxyapatite through fi'eeze drying”, Jom'nal of Materials Science: Materials in Medicine, 9, pp 583-587, (1998). S. Best, W. Bonfield, “Processing behaviour of hydroxyapatite powders with contrasting morphology”, Journal of Materials Science, pp 515-521, ( 1994). Champion E, Gautier S, BernacheAssolant D, “Characterization of hot pressed A1203 -p1ate1et reinforced hydroxyapatite composites”, Journal of Materials Science: Materials in Medicine, 7, pp 125-130, ( 1996). Yeong-Shyung Chou and David J. Green, “Silicon Carbide Platelet/Alumina composites: 11, Mechanical Properties”, Journal of the American Ceramic Society, 76, [6], pp 1452-1458, (1993). Tancret F, Desgardin G, Osterstock F, “Influence of porosity on the mechanical properties of cold isostatically pressed and sintered YBaQCu,O7_x superconductors”, Philosphical Magazine A, vol. 75 (2), pp 505-523 Feb. (1997). DR Clarke, “The Development of High-Tc Ceramic Superconductors: An Introduction”, Advanced Ceramic Materials, Volume 2/Number 3B/July 1987, Published by the American Ceramic Society (1987) Dale E. Grenoble, J. Lawrence Katz, Karl L. Dunn, Robert S. Gilmore and K. Linga Murty, “The Elastic Properties of Hard Tissues and Apatites”, Journal of Biomedical Research, Vol. 6 pp 221-233, (1972). 160 [81] [82] [831 [341 [85] [36] [87] [83] [39] [901 B.D. Case, J .R. Smyth, “Room-temperature fiacture energy of monoclinic Gd203”, Journal of Materials Science Letters 16 pp 3215-3217 (1981). Michel W. Barsoum, Fundamentals of Ceramics, McGraw-Hill 1997. Becher PF, “Microstructural design of toughened ceramics”, Journal of the American Ceramic Society, 74, [2], pp 255-269, (1991). ElRaghy T, Zavaliangos A, Barsoum MW, Kalidindi SR, “Damage mechanisms around hardness indentations in Ti3SiC2 ”, Journal of the American Ceramic Society, 80, [2], pp 513-516, (1997). R.W. Rice, “Limitations of pore-stress concentrations on the mechanical properties of porous materials”, Journal of Materials Science, 32, pp 4731-4736 (1997). Yi DQ, Li CH, “MoSiz-ZrO2 composites-fabrication, microstructure and properties”, Material Science and Engineering A Vol 261 :(1-2) pp 89-98 (1999). American Institute of Physics Handbook Second Edition, p 4-72, McGraw-Hill Book Company, Inc., New York, NY, 1963. Howard V. Malmstadt, Christie G. Enke, Stanley R. Church, Electronics and Instrumentation for Scientists, page 476, The Benjamin/Cummings Publishing Company, Inc., Reading MA, 1981. M.A. Fanovich, M.S. Castro, J.M. Porto Lopez, “Analysis of the microstructural evolution in hydroxyapatite ceramics by electrical characterisation”, Ceramics International, 25, pp 517-522, (1999). BR Lawn, A.G. Evans, “A model for crack initiation in elastic/plastic indentation fields”, Journal of Material Science, 12, pp 2195-2199, (1977). 161 1111 1 'lli‘llll IIIIIIIIIIIIIIIIIIIIIIIIIIIIIII 11[I[12111111111111111111