PROCESSING AND NANOSTRUCTURE INFLUENCES ON MECHANICAL PROPERTIES OF THERMOELECTRIC MATERIALS By Robert David Schmidt A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Materials Science and Engineering – Doctor of Philosophy 2014 ABSTRACT PROCESSING AND NANOSTRUCTURE INFLUENCES ON MECHANICAL PROPERTIES OF THERMOELECTRIC MATERIALS By Robert David Schmidt Thermoelectric (TE) materials are materials that can generate an electric current from a thermal gradient, with possible service in recovery of waste heat such as engine exhaust. Significant progress has been made in improving TE conversion efficiency, typically reported according to the figure of merit, ZT, with several recent papers publishing ZT values above 2. Furthermore, cost reductions may be made by the use of lower cost elements such as Mg, Si, Sn, Pb, Se and S in TE materials, while achieving ZT values between 1.3 and 1.8. To be used in a device, the thermoelectric material must be able to withstand the applied thermal and mechanical forces without failure. However, these materials are brittle, with low fracture toughness typically less than 1.5 MPa-m1/2, and often less than 0.5 MPa-m1/2. For comparison, window glass is approximately 0.75 MPa-m1/2. They have been optimized with nanoprecipitates, nanoparticles, doping, alterations in stoichiometry, powder processing and other techniques, all of which may alter the mechanical properties. In this study, the effect of SiC nanoparticle additions in Mg2Si, SnTe and Ag nanoparticle additions in the skutterudite Ba0.3Co4Sb12 on the elastic moduli, hardness and fracture toughness are measured. Large changes (~20%) in the elastic moduli in SnTe1+x as a function of x at 0 and 0.016 are shown. The effect on mechanical properties of doping and precipitates of CdS or ZnS in a PbS or PbSe matrix have been reported. Changes in sintering behavior of the skutterudite with the Ag nanoparticle additions were explored. Possible liquid phase sintering, with associated benefits in lower processing temperature, faster densification and lower cost, has been shown. A technique has been proposed for determining additional liquid phase sintering aids in other TE materials. The effects of porosity, grain size, powder processing method, and sintering method were explored with YbAl3 and Ba0.3Co4Sb12, with the porosity dependence of the elastic moduli reported. Only one other TE material has the porosity dependence of the elastic moduli previously reported in the literature, lead-antimony-silver-tellurium (LAST), and the effect of different powder processing and sintering methods has never been reported previously on TE materials. ACKNOWLEDGEMENTS I would like to acknowledge my advisor, Professor Eldon Case, for his patient advising and teaching throughout the work on this dissertation. Particular thanks to Professor Eldon D. Case for directing the projects and keeping everything moving, advising in all the experiments, and extensive assistance in advancing my writing. Thank you to Professor Timothy Hogan, Professor Donald Morelli and Professor Jeffrey Sakamoto for their work on this dissertation, the use of their labs for portions of the research, and for their work on my dissertation committee. I would like to acknowledge all my co-authors on the papers in this study, my lab mates, Jennifer Ni, Xiaofeng Fan, Cheng-Yun Liu, undergrad lab assistants Patricia Bordon Sarac, Miguel Valdes Tabernero, Andreia Ditzel Facci, Alex Zettler, Miguel Martinez Calderon, Luis Gonzalez Guerrero, Jesse Giles III, and summer high school honors student Zayra Lobo. Thank you to Karl Dersch for all of the PECS sintering of specimens. Thank you to the High Temperature Materials Laboratory at Oak Ridge National Laboratory. I would like to acknowledge the director, Edgar Lara-Curzio, for his support on the study of the SnTe thermoelectric materials, and making the resources of the laboratory available for use. I would like to acknowledge my primary contact for the user proposal, Rosa Trejo. Thank you to Professor Stanley Flegler, Carol Flegler and Abigail Vanderberg at the Center for Advanced Microscopy for their advice, maintenance and work with the scanning electron microscopes used in this dissertation. iv I would like to acknowledge the financial support of the Department of Energy, Revolutionary Materials for Solid State Energy Conversion Center, an Energy Frontiers Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under award number DE-SC0001054, for the funding for the research within this dissertation. v TABLE OF CONTENTS LIST OF TABLES ............................................................................................................ vii LIST OF FIGURES ........................................................................................................... xi 1 Introduction ...................................................................................................................... 1 1.1 Applications of thermoelectric materials .................................................................. 1 1.2 Thermoelectric efficiency ......................................................................................... 2 1.3 Mechanical properties of thermoelectric materials ................................................... 3 1.3.1 Elastic moduli .................................................................................................... 7 1.3.2 Fracture toughness ............................................................................................. 8 1.4 Porosity effects on mechanical properties ................................................................ 9 1.5 Temperature effects on mechanical properties ....................................................... 10 1.6 Sintering of thermoelectric materials ...................................................................... 11 REFERENCES ............................................................................................................. 12 2 Experimental Procedure ................................................................................................. 21 2.1 Material preparation ................................................................................................ 21 2.2 Crushing, grinding, sieving and regrinding (CGSR) process ................................. 22 2.3 Milling of CGSR powder ........................................................................................ 23 2.3.1 Planetary ball milling ....................................................................................... 23 2.3.2 Vibratory ball milling ...................................................................................... 24 2.4 Incorporation of nanoparticles ................................................................................ 26 2.5 Sintering of powders ............................................................................................... 26 2.5.1 Hot press sintering ........................................................................................... 26 2.5.2 Pulsed electric current sintering ....................................................................... 27 2.6 Sintered specimen preparation ................................................................................ 28 2.7 Mounting of specimens ........................................................................................... 28 2.7.1 Thermoplastic mounting of specimens ............................................................ 28 2.7.2 Thermoplastic dismounting of specimens ....................................................... 29 2.7.3 Epoxy mounting of specimens ......................................................................... 29 2.8 Specimen cutting ..................................................................................................... 30 2.9 Specimen polishing ................................................................................................. 30 2.10 Mass, dimensions and density............................................................................... 33 2.11 Resonant ultrasound spectroscopy ........................................................................ 33 2.12 Hardness and fracture toughness by Vickers indentation ..................................... 39 2.13 X-ray diffraction ................................................................................................... 42 2.14 Scanning electron microscopy imaging and energy dispersive x-ray spectroscopy ....................................................................................................................................... 42 2.14.1 Grain size measurement from fractured surface images ................................ 43 REFERENCES ............................................................................................................. 45 vi 3 Room temperature mechanical properties of polycrystalline YbAl3, a promising low temperature thermoelectric material ................................................................................. 48 Abstract ......................................................................................................................... 48 3.1 Introduction ............................................................................................................. 49 3.2 Experimental procedure .......................................................................................... 51 3.2.1 Materials and specimen preparation ................................................................ 51 3.2.2 Microstructural characterization ...................................................................... 54 3.2.3 Hardness testing ............................................................................................... 55 3.2.4 Elastic modulus measurements ........................................................................ 55 3.2.5 Fracture toughness measurements ................................................................... 56 3.3 Results and Discussion ........................................................................................... 57 3.3.1 Microstructural characterization ...................................................................... 57 3.3.2 Hardness ........................................................................................................... 62 3.3.3 Fracture Toughness .......................................................................................... 65 3.3.4 Elastic modulus ................................................................................................ 67 3.3.5. Acoustic wave speeds and the Debye temperature ......................................... 76 3.3.6 Mechanical properties as a function of processing technique ......................... 78 3.4 Summary and Conclusions ..................................................................................... 79 ACKNOWLEDGEMENTS .......................................................................................... 80 REFERENCES ............................................................................................................. 81 4 Room-Temperature Mechanical Properties and Slow Crack Growth Behavior of Mg2Si Thermoelectric Materials .................................................................................................. 86 Abstract ......................................................................................................................... 86 4.1 Introduction ............................................................................................................. 87 4.2 Experimental procedure .......................................................................................... 89 4.3 Results and Discussion ........................................................................................... 93 4.4 Conclusions ............................................................................................................. 99 ACKNOWLEDGEMENTS .......................................................................................... 99 REFERENCES ........................................................................................................... 101 5 Mechanical properties of Mg2Si thermoelectric materials with the addition of 0 to 4 volume percent silicon carbide nanoparticles (SiCNP) .................................................... 105 Abstract ....................................................................................................................... 105 5.1 Background ........................................................................................................... 105 5.2 Experimental Procedure ........................................................................................ 110 5.2.1 Materials and Specimen Preparation ............................................................. 110 5.2.2 Elasticity Measurements ................................................................................ 111 5.2.3 Hardness and toughness measurements ......................................................... 111 5.2.4 Microscopy .................................................................................................... 112 5.3 Results and Discussion ........................................................................................ 112 5.3.1 Microstructural analysis ................................................................................. 112 5.3.1.1 Starting material density and microstructure .......................................... 114 5.3.2 Elasticity results ............................................................................................. 117 5.3.2.1 Elasticity and SiCNP addition .................................................................. 117 vii 5.3.2.2 Elasticity and porosity............................................................................. 120 5.3.3 Hardness and Toughness Results ................................................................... 121 5.3.3.1 Crack bridging and toughness ................................................................. 126 5.4 Conclusions ........................................................................................................... 126 ACKNOWLEDGEMENTS ........................................................................................ 126 REFERENCES ........................................................................................................... 128 6 Influence of silver nanoparticle addition, porosity and processing technique on the mechanical properties of Ba0.3Co4Sb12 skutterudites ...................................................... 133 Abstract ....................................................................................................................... 133 6.1 Introduction ........................................................................................................... 134 6.2 Experimental procedure ........................................................................................ 137 6.2.1 Materials and Specimen Preparation ............................................................. 137 6.2.2 Elasticity Measurements ................................................................................ 138 6.2.3 Hardness and toughness measurements ......................................................... 140 6.2.4 Energy-dispersive X-ray spectroscopy and microscopy ................................ 141 6.3 Results and Discussion ......................................................................................... 149 6.3.1 Microstructural and chemical analysis ........................................................... 149 6.3.1.1 Microstructure of the AgNP and Ag agglomerates .................................. 149 6.3.1.2 Porosity and grain size and relationship to the processing technique ..... 152 6.3.2 Chemical analysis .......................................................................................... 153 6.3.2.1 Chemistry of the AgNP and Ag agglomerates ......................................... 153 6.3.2.2 Sintering behavior changes and scavenging of Sb .................................. 156 6.3.2.3 Possible significance to other thermoelectric material systems .............. 157 6.3.3 Elasticity results ............................................................................................. 158 6.3.3.1 Elasticity as a function of AgNP addition ................................................ 158 6.3.3.2 Elastic moduli and porosity .................................................................... 161 6.3.3.3 Elastic moduli and grain size/processing effects .................................... 164 6.3.4 Hardness and fracture toughness results ........................................................ 166 6.3.4.1 Hardness and toughness as a function of Ag addition ............................ 166 6.3.4.2 Hardness and fracture toughness as function of porosity ....................... 168 6.3.4.2.1 Hardness as function of porosity...................................................... 168 6.3.4.2.2 Fracture toughness as function of porosity ...................................... 169 6.3.4.3 Hardness and fracture toughness as a function of load ........................... 170 6.3.4.4 Hardness and fracture toughness results as a function of grain size ....... 172 6.3.4.4.1 Hardness results as a function of grain size ..................................... 172 6.3.4.4.2 Fracture toughness results as a function of grain size...................... 173 6.4 Summary and conclusions .................................................................................... 175 ACKNOWLEDGEMENTS ........................................................................................ 176 REFERENCES ........................................................................................................... 177 7 High temperature elastic moduli of thermoelectric SnTe1+X –y SiC nanoparticulate composites....................................................................................................................... 187 Abstract ....................................................................................................................... 187 7.1 Introduction ........................................................................................................... 188 viii 7.2 Experimental procedure ........................................................................................ 190 7.2.1 Specimen Preparation .................................................................................... 190 7.2.2 Specimen characterization ............................................................................. 192 7.2.3 High temperature Resonant Ultrasound Spectroscopy measurements .......... 193 7.3 Results and Discussion ......................................................................................... 194 7.3.1 Microstructural analysis ................................................................................. 194 7.3.2 Elastic modulus as a function of temperature, matrix composition and SiCNP volume fraction ....................................................................................................... 201 7.3.2.1 Bilinear model of elastic modulus versus temperature ........................... 205 7.3.2.2 Effects of chemical composition and porosity on elastic moduli of the SnTe1+X matrix .................................................................................................... 210 7.3.2.3 Comparison of the Young’s modulus versus elasticity behavior of polycrystalline SnTe1+X to the literature ............................................................. 215 7.3.2.4 Effect of volume fraction of SiCNP on the elastic moduli of the SnTe1+X SiCNP composites ................................................................................................ 217 7.5 Summary and conclusions .................................................................................... 220 ACKNOWLEDGEMENTS ........................................................................................ 221 REFERENCES ........................................................................................................... 222 8 Mechanical properties of lower-cost, earth-abundant chalcogenide thermoelectric materials, PbSe and PbS, with additions of 0 to 4% CdS or ZnS ................................... 228 Abstract ....................................................................................................................... 228 8.1 Introduction ........................................................................................................... 229 8.2 Experimental procedure ........................................................................................ 231 8.2.1 Materials and Specimen Preparation ............................................................. 231 8.2.2 Elasticity Measurements ................................................................................ 232 8.2.3 Hardness and toughness measurements ......................................................... 234 8.2.4 Microscopy and X-ray diffraction ................................................................. 234 8.3 Results and Discussion ......................................................................................... 235 8.3.1 Microstructural analysis ................................................................................. 235 8.3.2 Elasticity results ............................................................................................. 245 8.3.3 Hardness results ............................................................................................. 248 8.4 Discussion ............................................................................................................. 248 8.4.1 Microstructural analysis ................................................................................. 248 8.4.2 Elasticity analysis........................................................................................... 250 8.4.3 Hardness analysis ........................................................................................... 254 8.5 Conclusions ........................................................................................................... 257 ACKNOWLEDGEMENTS ........................................................................................ 258 REFERENCES ........................................................................................................... 259 9 Summary and Conclusions .......................................................................................... 267 9.1 Porosity dependence and processing method independence of elastic moduli..... 267 9.1.1 Grain size and processing method independence of elastic moduli .............. 268 9.2 Sintering aids and TE materials ............................................................................ 268 9.3 Elastic moduli of TE materials with nanoparticle or nanoprecipitate additions ... 269 ix 9.4 Fracture toughness and slow crack growth in Mg2Si and other TE materials ...... 270 REFERENCES ........................................................................................................... 272 10 Future work ................................................................................................................ 274 REFERENCES ........................................................................................................... 276 APPENDICES ................................................................................................................ 278 APPENDIX A. Effect of nanoparticle addition on the elastic modulus of a composite material ....................................................................................................................... 279 APPENDIX B. Standard operating procedure for HT-RUS furnace .......................... 283 APPENDIX C. Machine drawings of the HT-RUS equipment .................................. 303 REFERENCES ........................................................................................................... 362 x LIST OF TABLES Table 2.1 ........................................................................................................................ 32 A series of diamond grit pastes were used to polish the specimens. A grit size may be skipped, with additional polishing time on the next smaller diamond grit size to compensate. Table 3.1 ........................................................................................................................ 53 The HP and PECS processing parameters for the YbAl3 specimens included in this study. Table 3.2 ........................................................................................................................ 58 Specimen label, densification method (hot pressed, HP, or pulsed electric current sintering, PECS), mass, dimensions and mass density of each of the YbAl3 specimens included in this study. Table 3.3 ........................................................................................................................ 58 The number of indentations, indentation load and mean hardness, , of the YbAl3 specimens and porosity, P. In each case, the loading time was 10 s. The porosity reported is the average as measured from the disc and bar specimen geometries. Table 3.4 ........................................................................................................................ 66 For YbAl3 specimens included in this study, the fracture toughness, KC, measured by Vickers indentation, including the applied load, specimen grain size and volume fraction porosity, P. The KC was obtained from the mean of 10 indentations per specimen. Table 3.5 ........................................................................................................................ 69 The fitting parameters obtained from the least-squares fit of the appropriate elastic modulus to Eqs. (5a) – (5c). The coefficient of determination, R2, is given for each of the fitted equations. Table 3.6 ........................................................................................................................ 69 A comparison of this study’s experimentally determined values of elastic moduli (E, Young’s modulus, G, shear modulus, B, bulk modulus and , Poisson’s ratio) for YbAl3 with calculated and experimental values of other trialuminide materials, R-Al3, from the literature for R = Yb, Sc, Tm, or Zr [Sa 2011, Zhou 2010, Tao 2008, Hyland 1991, Fu 1990, Xu 1990, Jahnatek 2007]. xi Table 3.7 ....................................................................................................................... 75 For the seven YbAl3 specimens included in this study, N, the number of RUS resonant peaks measured, the RMS error in RUS, the Young’s modulus, E, shear modulus, G, Poisson’s ratio, ν, longitudinal and shear velocity, VL and VS, acoustic Debye temperature, θD, and porosity, P, were measured. The RUS analysis on each specimen was performed on the tripod configuration (Fig. 3.1a). Table 3.8 ........................................................................................................................ 77 The fitting parameters obtained from the least-squares fit of the VL and VS versus P data to Eqs. (6a) and (6b), respectively, with the coefficient of determination, R2. Table 4.1 ........................................................................................................................ 88 The Mg2Si specimens used in this study for Vicker’s hardness, H, fracture toughness, KIC, Young’s modulus, E, shear modulus, G, Poisson’s ratio, , and acoustic debye temperature, D, measurements. The Mg2Si pieces were reduced using a (1) mechanical mortar and pestle until all powder passed through a 53 µm sieve and then either densified or further reduced using a (2) planetary mill at 150 rpm for 3 hours. For indentation the specimens were polished to 1 μm diamond grit, further details are available elsewhere [Ni 2010]. Due to reactions with water [Aesar], the specimens were cleaned using ethanol instead of water between polishing grit sizes. Table 4.2 ........................................................................................................................ 92 The density, Young’s modulus, E, shear modulus, G, Poisson’s ratio, , and Debye temperature, D, from this study compared with the literature. The moduli were determined using resonant ultrasound spectroscopy, RUS, first principle calculation, FPC, resonance technique, RT, and compression. Table 4.3 ........................................................................................................................ 95 A comparisons of Vickers hardness, H, and fracture toughness, KIC, from this study to the literature. For the PECS processed specimens in this study, the H and KIC was determined using at least 10 indentations per load with one standard deviation reported as error. For cast specimens in literature, Milekhine et al. measured at least 10 indentations per load [Milekhine 2002]and Li et al. measured 5 indentations [Li 1993]. xii Table 4.4 ........................................................................................................................ 98 No significant slow crack growth occurred for the six Vickers indentation cracks listed below for specimen Mg2Si-01, where the radial crack lengths were monitored for up to 3 days (Fig. 4.5). For the radial crack length measurements performed over the entire time range, the measured crack lengths did not depart significantly from the mean length of each crack, as indicated by the coefficient of variation (CV, standard deviation/mean), where CV ranges from about 1.4% to 5%. Table 5.1 ........................................................................................................................ 109 The fracture toughness, KC, has been improved in brittle materials, including the thermoelectric Bi2Te3, by the addition of SiC nanoparticles. Table 5.2 ........................................................................................................................ 109 Doped Mg2Si-based thermoelectric materials with ZT near 1 have been reported. These reasonable ZT values for a thermoelectric material support the use of Mg2Si as a model system. Table 5.3 ........................................................................................................................ 115 Specimens in this study were either milled by planetary ball mill (PM) or vibratory mill (VM), with up to 4 vol% SiCNP additions, then sintered by pulsed electric current sintering to produce specimens with a density (ρ) of 2.00 g/cm3 or greater for all specimens except VM-0SiC-2. The average grain size (GS) by the lineal intercept method are a function of the milling method, and not a function of SiCNP additions. Table 5.4 ........................................................................................................................ 119 The Young’s modulus, E, shear modulus, G, and density, ρ, for the specimens in this study, as compared to the range of E, G and ρ in a previous study on three Mg2Si specimens produced by the same vendor. Table 6.1 ........................................................................................................................ 139 Specimens of skutterudite Ba0.3Co4Sb12 used in this study varied based on AgNP addition, sintering procedure and temperature, resulting porosity, and powder processing. Specimens VM-PECS-673, VMPECS-723 and VM-PECS-773 were reprocessed from specimens PMHP-673, PM-HP-773 and PM-HP-873 by grinding and powder processing by vibratory milling, and sintering by PECS. Table 6.2 ........................................................................................................................ 154 Results of EDS scan using Point ID mode on two as-received AgNP agglomerates. xiii Table 6.3 ........................................................................................................................ 154 Results of EDS scan using Point ID mode on specimen PM-HP-873-Ag. Spacing between spots is approximately 5 to 10 µm (Figure 6.4e). Ag areas examined are 1-3 µm in diameter. Table 6.4 ........................................................................................................................ 154 Results of EDS scan using Point ID mode on specimen PM-HP-773-Ag. Spacing between spots is approximately 4 µm (Figure 6.4c). Ag area examined approximately 1-3 µm in diameter. Table 6.5 ........................................................................................................................ 154 Results of EDS scan using Point ID mode on specimen PM-HP-673-Ag. Spacing between spots is approximately 5 to 10 µm (Figure 6.4a). Ag area examined is approximately 20 µm in diameter. Table 6.6 ........................................................................................................................ 160 The linear decrease in elastic moduli, E, G, and B, according to equation (5b), for this study of Ba0.3Co4Sb12 is consistent with the limited information available in the literature for porosity dependence of elastic moduli for thermoelectric materials [Schmidt 2013a; Ni 2009]. Table 6.7 ........................................................................................................................ 165 For CoSb3-based thermoelectric materials, a comparison of the KC results from this study with the literature, including the porosity, P, the grain size, GS, of the specimens and the KC measurement technique [Ravi 2009; Eilertsen 2013; Rogl 2011]. Table 7.1 ........................................................................................................................ 191 For each of the six SnTe-SiCNP specimens included in this study, the volume fraction SiCNP addition, specimen mass, dimensions, mass density, porosity, P, and mean grain size. Specimens were prepared from one of two starting ingots, designated A or B in the specimen label. Table 7.2 ........................................................................................................................ 200 For SiCNP clusters identified by electron backscatter images for specimen SnTe(x = 0)V0.04 (Fig. 6), the SEM magnification, the area of the field of view of the micrograph, along with the number, number density, and size range of the SiCNP clusters Table 7.3 ........................................................................................................................ 203 For the six SnTe-SiCNP specimens included in this study, results of a least squares fit to equation (4a) for the Young’s modulus, E, versus temperature, T, data and to equation (4b) for the shear modulus, G, versus T, data. The coefficients of determination for the fit of the E versus T and the G versus T data are given by and , respectively xiv Table 7.4 ........................................................................................................................ 206 For two separate temperature ranges, RT – 543 K, and 543 K – 663 K, the E versus T was fit to equation (4a). Table 7.5 ........................................................................................................................ 206 Comparison of Δb, equation (6), for this study and LAST [Ren 2010]. Table 7.6 ........................................................................................................................ 211 Fractional changes in Young’s modulus, E, shear modulus, G, and hardness H as a function of changes in composition from n1 to n2 [this study, [Kawaharada 2004; Ravinder 2001; Schenk 1998]. Table 7.7 ........................................................................................................................ 214 In order to compare the room temperature experimental values of Young’s modulus, Eexp, and shear modulus, Gexp, for this study’s specimens without SiCNP additions to the single crystal values of E0 and G0 values from SnTe from the literature [Beattie 1969; Simmons 1971], equations (7c) and (7d) were used to calculate porosity corrected values for the single crystal data using the modulus-porosity slope data for two thermoelectric materials, namely YbAl3 (bPE = 2.34, bPG = 2.38) and LAST (bPE = 3.5, bPG = 3.5) [Schmidt 2013; Ni 2009]. This study’s experimental Eexp and Gexp values for the tellurium rich specimen, SnTe(x=0.016)V0, are in the best agreement with the range of porosity corrected values ECORR and GCORR. Table 7.8 ........................................................................................................................ 216 Comparison of the Young’s modulus, E, versus temperature behavior for a variety of brittle materials, SnTe [this study], PbTe [Houston 1968; Simmons 1971; Ren 2010], and selected skutterudite TE materials [Schmidt 2012; Ravi 2008]. The parameters ERT and bRT were obtained via a least-squares fit of each of the data sets to equation (4a). The coefficient of determination, R2, was equal to or greater than 0.98 for each data set, indicating that equation (4a) describes the E versus temperature behavior well. Table 8.1 ........................................................................................................................ 236 Composition, theoretical density, theo, measured density, meas, volume fraction porosity, P, average grain size, , typical observed inclusion size range, Incl, and mode mixity, M (equation 3) for the PbSbased specimens included in this study. xv Table 8.2 ........................................................................................................................ 237 Composition, theoretical density, theo, measured density, meas, volume fraction porosity, P, average grain size, , typical observed inclusion size range, Incl, and mode mixity, M (equation 3) for the PbSebased specimens included in this study. Table 8.3 ........................................................................................................................ 244 EDS results from four area scans of a polished PbS specimen, indicating a higher concentration of Pb than S. Table 8.4 ........................................................................................................................ 252 The Young’s modulus, E, the shear modulus, G, and the Poisson’s ratio measured in this study for the polycrystalline undoped PbSe and undoped PbS specimens compared with the aggregate average values of E and G computed from the Hashin and Shtrikman bounds, [Simmons 1971] for single crystal elasticity from the literature [Dalven 1969, Bhagavantam 1951, Lippmann 1971, Hellwege 1979]. Table 8.5 ........................................................................................................................ 255 In general, hardness, H, is grain size dependent, and the single crystal hardness results from loem and r ger [Bloem 1955] are expected to have a lower hardness than the other, polycrystalline specimens in this table. The grain size for the PbS and PbSe was not listed by Darrow [Darrow 1969]. Average grain sizes for all specimens in this study are between 1.8 µm and 18.7 µm, listed in Tables 8.1 and 8.2. Table A.1........................................................................................................................ 255 For spherical metallic particle additions to brittle matrices [Fujieda 2012; Hasselman 1965], comparison of the measured composite modulus, EC, with the values predicted by the four models given in this appendix (rule of mixtures, ROM, Reuss constant strain, RCS, Hashin particulate, HP, and Halpin-Tsai, HT [Schmidt 2013b]). Em is the Young’s modulus of the matrix material [Fujieda 2012; Hasselman 1965]and Er is the Young’s modulus of the reinforcing material [Lowrie 1967; Macfarlane 1965; Neighbours 1958; Chang 1966]. xvi LIST OF FIGURES Figure 1.1 ....................................................................................................................... 6 Schematic of a thermoelectric couple, indicating current flow when placed in a temperature gradient. Figure 2.1 ....................................................................................................................... 36 Typical setup of tripod arrangement of RUS transducers, with specimen on top and the transducer tips near the edge of the specimen. Figure 2.2 ....................................................................................................................... 36 RUS spectrum from PbSe specimen. Each sharp peak in intensity represents a mechanical resonance at the driven frequency. Figure 2.3 ....................................................................................................................... 37 RUS analysis software, showing setup of initial conditions for a PbSe specimen. Figure 2.4 ....................................................................................................................... 38 Sample output file from a CylModel RUS analysis, truncated to show only the first 12 resonance peaks. Note that resonance frequency 10, 191.681 kHz, was not included in the analysis to improve the error. The “chisquare increased 2% by…” section includes the highlighted values of A, 0.47, and of D, 0.11, used in the determination of the uncertainties, dc11 and dc44. Figure 2.5 ....................................................................................................................... 41 Schematic representation of Vickers indentation, illustrating the indentation impression dimensions, 2a, and the radial crack length dimensions, 2c. Figure 3.1 ....................................................................................................................... 58 Photographs of tripod and bipod transducer stages for the RUS measurements of elastic modulus performed in this study. Figure 3.2 ....................................................................................................................... 59 SEM micrographs of fracture surfaces and polished surfaces of YbAl3 specimens included in this study. The ovals indicate more porous regions neighboring more dense areas that occur in the two HP specimens (Fig. 3.2a and b) as well as the PECS specimen intentionally processed to give high porosity (Fig. 3.2e). xvii Figure 3.3 ....................................................................................................................... 60 The fractured surfaces of specimens YbAl3-B and YbAl3-E exhibited lower porosity regions surrounded by channels of higher porosity in (a) and (b), both with similar porosity but each sintered by a different method. In contrast, specimen YbAl3-D, shown in (c), exhibited a uniform density throughout. Figure 3.4 ....................................................................................................................... 64 The hardness, H, versus volume fraction porosity, P, for the YbAl3 specimens included in this study. The solid line represents a leastsquares fit to Eq. (4b). Figure 3.5 ....................................................................................................................... 70 For the YbAl3 specimens included in this study, the RUS measurement results for (a) the Young’s modulus, E, shear modulus, G, bulk modulus, B as a function of P and (b) Poisson’s ratio as a function of P. The solid lines in Fig. 3.5a represent least-squares fits to Eqs. (5a), (5b) and (5c) for the Young’s modulus, shear modulus and bulk modulus, respectively. Figure 3.6 ....................................................................................................................... 71 The RUS mechanical vibration spectra for the two-transducer stage and tripod specimen stage. Figure 3.7 ....................................................................................................................... 77 From the RUS measurements of the YbAl3 specimens included in this study, (a) shear and longitudinal wave speeds as a function of P, where the solid lines in represent least-squares fits to Eq. (6a) and (6b) for the longitudinal and shear velocities, respectively. (b) the effective Debye temperature as a function of P, where the solid line represents a leastsquares fit to Eq. (9). Figure 4.1 ....................................................................................................................... 88 The time–temperature and time–pressure profiles used in the PECS processing of each of the Mg2Si specimens included in this study. Figure 4.2 ....................................................................................................................... 91 X-ray diffraction of planetary-milled Mg2Si powder with peaks compared with the literature for MgO [Hazen 1976] and Mg2Si [Owen 1923]. xviii Figure 4.3 ....................................................................................................................... 91 Fractured surface of specimens (a) Mg2Si-01 and (b) Mg2Si-05, PECS processed from planetary mill and mechanical mortar and pestle Mg2Si powder, respectively. Cleavage steps show mixed- mode fracture, with both intergranular and transgranular fracture present. Entrapped pores visible at grain boundaries were approximately 1 µm to 3 µm diameter with polygonal shape. The dark areas between grains are porosity within the specimens, typically around 1 µm or less in diameter. Figure 4.4 ....................................................................................................................... 92 Typical Vickers indentation crack systems for specimens Mg2Si-01 and Mg2Si-05, which display indentation impressions, little chipping or spalling, and a fully developed radial crack system. Figure 4.5 ....................................................................................................................... 97 For specimen Mg2Si-01, plots of radial crack length versus time for six Vickers indentation crack systems loaded at 2.94 N. In figures (b) and (e), a radial crack popped in within 1 min to 5 min following the initial indentation event. The average crack length and one standard deviation variation is indicated by solid and dashed lines, respectively, for each indentation crack. The crack length may vary significantly between indentations, resulting in the uncertainties in the reported KIC values in Table 4.3. Figure 5.1 ....................................................................................................................... 113 Planetary milled Mg2Si powder exhibited typical particle sizes of submicron to 5 μm in SEM. The powder has a surface area of 4.4 m2/g measured by ET, or approximately 0.7 μm average particle size. Figure 5.2 ....................................................................................................................... 113 Vibratory milled Mg2Si powder exhibited typical particle sizes of 0.2 μm to 2 μm in SEM. Figure 5.3 ....................................................................................................................... 116 For Mg2Si fracture surfaces, transgranular fracture dominate in all specimens. The PM specimens (a-c) with varying amounts of SiCNP addition did not show any appreciable difference in grain size, although changing to VM processing significantly reduced the grain size (d). Note the difference in scale between the PM images (a-c) and the VM image (d). xix Figure 5.4 ....................................................................................................................... 118 The Young’s modulus (a, b) and shear modulus (c, d) of Mg2Si varied primarily due to small variations in density. The variation in vol. % SiCNP did not significantly affect the moduli, regardless of if the specimens were planetary milled (a, c) or vibratory milled (b, d). Figure 5.5 ....................................................................................................................... 118 The Young’s modulus, E, decreases linearly with porosity for the set of 13 specimens in this study. Figure 5.6 ....................................................................................................................... 122 The hardness, (a) and (b) is not a function of the milling procedure or the vol% SiCNP, but less scatter was observed in the (b) vibratory milled specimens than the (a) planetary ball milled specimens. The fracture toughness exhibited a maximum at 1 vol% SiCNP for the (c) planetary ball milled specimens, but the fracture toughness is not a function of vol% SiCNP for the (d) vibratory milled specimens. Open symbols in (b) and (d) indicate a specimen with lower density of 1.93 g/cm3, relative to the 2.00 to 2.06 g/cm3 for all other specimens in this study. Figure 5.7 ....................................................................................................................... 124 Crack bridging, in PM Mg2Si was commonly observed in radial cracks for all the Mg2Si specimens in this study. Figure 5.8 ....................................................................................................................... 125 Crack bridges were commonly observed in VM specimens regardless of SiCNP addition. Crack bridging in radial cracks was not eliminated by reducing grain size through VM processing. Spotting is from oil residue on the surface of the specimen. Figure 6.1 ....................................................................................................................... 139 RUS spectrum from specimen PM-HP-873-Ag, P = 0.05 with AgNP. The elastic moduli of each of the specimens in this study are calculated from the specimen mass, dimensions, shape, and resonant frequencies. Each peak in the spectrum represents a mechanical resonance at that frequency Figure 6.2 ....................................................................................................................... 142 Silver nanoparticles exhibited agglomerates of 10 µm or greater (a-c), but consisting of individual grains or particles consistent with the manufacturer claimed average particle size of 20-40 nm (d) xx Figure 6.3 ....................................................................................................................... 143 Polished surfaces of specimen PM-HP-873-Ag, both in secondary electron images (a) and backscatter images (b). Porosity is observed between larger grains (a and b), but also with areas of Ag visible in backscatter as bright areas (b), due to the higher average atomic weight of the silver rich regions. The abundance of Ag and the relative deficiency of Co or Sb in the bright areas of the backscatter image (b) is confirmed by energy-dispersive x-ray spectroscopy maps (c-e) Figure 6.4 ....................................................................................................................... 144 For each of the hot pressed specimens with AgNP additions, an EDS line scan was performed and point ID locations were chosen. For PM-HP673-Ag, a cut surface was used (a) because the specimen was not successfully polished. For PM-HP-773-Ag and PM-HP-873-Ag, polished surfaces were examined (c and e). The line scan for PM-HP673-Ag (b) indicated only silver present except for two regions in the silver where antimony and cobalt were present in concentrations consistent with skutterudite particles. In contrast, the line scan for PMHP-773-Ag (d) and PM-HP-873-Ag (f) indicated the presence of antimony in the silver-rich locations. No silver was observed in the matrix outside of the silver-rich locations Figure 6.5 ....................................................................................................................... 145 The microstructure of Ba0.3Co4Sb12, without Ag addition, changed as a function of sintering temperature. For (a) a sintering temperature of 673 K, microstructure and porosity, P, are consistent with a specimen with little to no observed sintering or densification. After sintering at 773 K (b), only minimal densification and neck formation are observed. Hot pressing at 873 K (c) enhanced both grain growth and densification in the specimen without Ag nanoparticles, however, the microstructure still is observed to be very porous, consistent with measurement of P = 0.16 Figure 6.6 ....................................................................................................................... 146 The microstructure of Ba0.3Co4Sb12 with 0.5 wt% AgNP addition changed as a function of sintering temperature. At 673 K (a), little to no sintering is observed in the microstructure. Hot pressing at 773 K (b) resulted in limited grain growth and densification, with necks observed between the grains but significant porosity observed, consistent with early stage sintering. The porosity for the specimen with Ag nanoparticles sintered at 873 K (c) is the lowest (P = 0.03) of all the specimens in this study xxi Figure 6.7 ....................................................................................................................... 147 Fracture surface images of reprocessed material sintered by PECS exhibit dense regions surrounded by regions of higher porosity. All regions of the specimens exhibit similar grain sizes. The porous areas decrease as the sintering temperature increased from (a) 673 K, to (b) 723 K, to 773 K Figure 6.8 ....................................................................................................................... 148 After reprocessing by SPEX milling and sintering by PECS, the samples exhibit a unimodal, sub-micron grain size distribution. Figure 6.9 ....................................................................................................................... 151 Hand milling of AgNP agglomerates in ethanol were not observed to reduce the size of the nanoparticle agglomerates. After hand milling, silver nanoparticle agglomerates of 10 µm or greater were typically observed, similar to the size of agglomerates for the as received AgNP. Several faces of the agglomerates appear to be deformed after hand milling Figure 6.10 ..................................................................................................................... 159 The (a) Young’s modulus, E, (b) shear modulus, G, and (c) bulk modulus, B, are each a function of porosity. In each figure, the solid lines represent a least-squares fit to equation (5b). The Poisson’s ratio (d) is observed to be a weak function of porosity. The elastic moduli were not observed to be a function of the addition of 0.5 wt% AgNP or of reprocessing Figure 6.11 ..................................................................................................................... 163 The elastic moduli for the exact composition of skutterudite in this study are not recorded in literature, however the moduli for specimens of similar composition [Zhang 2010] are consistent with the porosity dependent elastic moduli relationships observed in this study. Filled symbols indicate specimens with 0.5 wt% AgNP, unfilled for specimens without any AgNP, and half-filled symbols for reprocessed specimens without any AgNP. The solid lines represent a least squares fit to equation (5b) of the data in this study xxii Figure 6.12 ..................................................................................................................... 167 The (a) hardness and (b) fracture toughness for the hot pressed specimens with AgNP, filled symbols, the hot pressed specimens without AgNP, open symbols, and the reprocessed PECS specimens without AgNP, half-filled symbols, from Vickers indentation at four loads, 9.8 N, 4.9 N, 2.94 N, and 1.96 N. Full radial cracks were not observed on the specimens at 1.96 N load and therefore no fracture toughness values are available at 1.96 N load. The hot pressed specimens sintered at 673 K were not able to be polished and were not tested. The solid lines in the figure (a) represent hardness for the reprocessed specimens were fit to equation (6) separately from the hot pressed specimens because hardness is a function of grain size. Fracture toughness, figure (b), was not observed to be a function of either porosity or load, as plotted by the average (solid) and standard deviation (dotted) lines, with an average KC for all 7 specimens of 1.0 ± 0.2 MPa-m½ Figure 6.13 ..................................................................................................................... 171 For each of the specimens, a plot of c3/2 versus load, F, was used to determine the suitability of the fracture toughness by Vickers indentation. For 5 of 7 specimens, the coefficient of determination, R2, for a linear regression to equation (7) through the data points is 0.99 or greater, with an R2 of 0.98 for specimen VM-PECS-773, and 0.84 for specimen PM-HP-773. Note specimen PM-HP-773 also has a low fracture toughness measurement at 2.94 N load. All specimens were measured at three loads, except VM-PECS-673 and VM-PECS-773 with a fourth measurement at a Vickers indentation load of 1.96 N Figure 7.1 ....................................................................................................................... 191 Processing parameters (temperature and die pressure) as a function of time for the PECS densification of the SnTe-SiCNP specimens included in this study. Figure 7.2 ....................................................................................................................... 195 For SnTe specimen A-06 (0 vol% SiCNP), (a) fractured specimen surface and (b) polished specimen surface. Figure 7.3 ....................................................................................................................... 195 The as-received SiCNP consisted of particles roughly 50 nm across agglomerated into clusters ranging in size from approximately 100 nm to 20 µm. xxiii Figure 7.4 ....................................................................................................................... 195 (a) After grinding in an alumina mortar and pestle for 5 min in ethanol, the SiCNP exhibited agglomerations of particles ranging from sub-micron to ~10 µm diameter. (b) The agglomerations consisted primarily of nanoparticles with diameter less than 100 nm, but with an occasional micron-scale particle. The size distribution of the agglomerates after manual grinding was similar to that observed prior to manual grinding (Figure 7.3). Figure 7.5 ....................................................................................................................... 197 Fractured specimen surface images for (a) specimen SnTe (x = 0)V0 with no SiCNP added, and for (b) SnTe (x = 0)V0.01, (c) SnTe(x = 0)V0.02, (d) SnTe(x = 0)V0.03, and (e) SnTe(x = 0)V0.04. In (b) – (e), the location of SiCNP clusters is indicated by arrows. Figure 7.6 ....................................................................................................................... 199 Specimen SnTe(x = 0)V0, with no added SiCNP, shows no surface features other than surface debris in both (a) a secondary electron image (SEI) and in (b) a backscattered electron image. Specimen SnTe(x = 0)V0.04 with 0.04 vol fraction SiCNP addition shows a difference between the (c) secondary electron image and (d) the backscattered electron image, with the image in the backscattered electron mode showing dark regions indicating likely SiCNP clusters at or near the specimen surface. Figure 7.7 ....................................................................................................................... 200 A backscattered electron image of a polished surface of specimen SnTe(x = 0)V0.04 indicates darker regions with a lower average atomic weight, likely composed primarily of the SiCNP. Using EDS, silicon was detected at the location of spectrum one, at approximately 13 at%, but not at spectrum two, with <1 at%, confirming that the dark regions contain a high concentration of the SiCNP. The specimen was osmium coated prior to imaging to reduce localized charging. Figure 7.8 ....................................................................................................................... 202 The Young’s modulus versus temperature for the six SnTe–SiCNP specimens included in this study. For each specimen, there is no observable hysteresis between the heating and cooling curves indicating the lack of significant microcracking or bloating over the test temperature range (room temperature to 663 K). The error bars are smaller than the plotting symbols for each modulus value in (a–f). xxiv Figure 7.9 ....................................................................................................................... 203 Schematic representation of the bilinear Young’s modulus, E, versus temperature, T, behavior showing TKINK and the associated slope change in the E versus T curve. Figure 7.10 ..................................................................................................................... 204 The residuals (equation 5) representing the difference between the experimental Young’s modulus data and the modulus values predicted from the least-squares fit of the E versus T data to equation (4). The dashed lines at 0.05 GPa are a guide, representing roughly ± 0.01 ERT, where ERT is defined as the room temperature intercept of the modulus– porosity relationship given by equation (4a). A change in the E versus T slope occurs at approximately 543 K (TKINK). Figure 7.11 ..................................................................................................................... 208 Unlike Figure 7.10, the residuals represent the difference between the experimental data and the least-squares fit performed to equation (4) in a piecewise manner, that is least-squares fits were performed separately for the two temperature intervals (i) TRT to TKINK and (ii) TKINK to the maximum test temperature. As in Figure 7.10, the dashed lines at 0.05 GPa represent roughly ± 0.01 ERT, where ERT is defined as the room temperature intercept of the modulus-porosity relationship given by equation (4a). In (a) – (f), the vertical line at 543 K represents TKINK, the temperature at which the change in the E versus T slope occurs. Figure 7.12 ..................................................................................................................... 219 Comparison of the experimental Young’s modulus, E, results for the composites SnTe1+X –ySiC (y = 0.0 to 0.04) with the four composite models given by equations (8) – (12). Also plotted are the values of E corrected to zero porosity using equation (7a). Figure 8.1 ....................................................................................................................... 233 RUS scan of PbSe specimen Figure 8.2 ....................................................................................................................... 238 Fracture surfaces of PbS specimens indicate primarily transgranular fracture in undoped PbS (a) changing to intergranular fracture for the Na-doped PbS specimen (b). With 1% to 4% addition of CdS, the fracture mode changes to primarily transgranular fracture (b, d), and with 1% to 4% addition of ZnS, the fracture mode changes to mixed, with majority transgranular and minority intergranular (c, e). Note the bright areas in images (c) and (e) are the ZnS precipitates. xxv Figure 8.3 ....................................................................................................................... 239 Fracture surfaces of PbSe specimens indicate primarily intergranular fracture for the pure PbSe specimen, the Na-doped PbSe specimen, and the specimens with 1% addition of CdS or ZnS (a-c). With 2% to 4% addition of CdS or ZnS, the fracture mode changes to primarily transgranular fracture (d-f), although significant intergranular fracture was also observed in the specimen with 4% ZnS (f). Note the bright areas in images (d) and (f) are the ZnS precipitates Figure 8.4 ....................................................................................................................... 240 Secondary images of the polished specimens show isolated, micron scale spherical porosity and minor scratches from polishing. In the specimens with 4% CdS or ZnS, precipitates of sub-micron up to approximately 15 μm were observed, particularly in backscatter electron (BSE) mode. The images of polished PbS in both secondary electron (SE) and BSE mode (a and b) only exhibit spherical pores ~1 µm, with no evidence of precipitates, consistent with the material having no additions of CdS or ZnS. Figure 8.5 ....................................................................................................................... 242 Observed in backscatter, precipitates of CdS were up to ~4 μm, with some micron scale precipitates with geometry of rods or plates and sharp facets consistent with crystallographic alignment. Figure 8.6 ....................................................................................................................... 242 Some ZnS inclusions in PbS resembled stacked plates. Figure 8.7 ....................................................................................................................... 244 XRD pattern of the sintered PbS specimen. Figure 8.8 ....................................................................................................................... 246 The E and G of the undoped PbS and PbSe were higher than the Nadoped PbS and PbSe. The E and G of undoped PbS are likely higher due to reduced porosity (pure PbS P = 0.008, all other PbS P = 0.031 to 0.050), while the E and G of undoped PbSe are likely higher due to doping effects. Only small changes in E and G were noted in Na-doped PbS and PbSe with addition of up to 4% CdS or ZnS, with. Only small variability in the ν noted in PbS with addition of CdS or ZnS. The ν of PbSe increased with the addition of CdS or ZnS. Solid line is average and dotted line is standard deviation of the specimens with 1% to 4% CdS or ZnS addition. xxvi Figure 8.9 ....................................................................................................................... 247 No measurable change in hardness was noted in PbS with addition of up to 4% CdS or ZnS. Hardness of PbSe increased from 0.5 GPa to 0.8 GPa with the addition of either CdS or ZnS. The scatter in the fracture toughness data of PbS increased with increasing CdS or ZnS content. No change in the fracture toughness of PbSe was noted. Figure B.1. ..................................................................................................................... 284 The top of the furnace has exposed leads to power the heating elements with up to 120 VAC electricity. Do not operate the furnace with these leads exposed. Figure B.2. ..................................................................................................................... 285 Note feedthroughs for the thermocouples (left arrow) and heating element power (right arrow). These feedthroughs and the wires from them should not touch the base or the sides of the bell jar when closed. Figure B.3. ..................................................................................................................... 285 Braided insulation must be in place around the conductors to the heating elements to prevent a short. Figure B.4. ..................................................................................................................... 286 The o-ring and the base of the bell jar may have some debris, particularly from the furnace refractories. The debris is abrasive, and may inhibit a proper seal between the bottom and side of the bell jar. The debris should be removed and appropriate grease reapplied to the o-ring prior to operation. Figure B.5. ..................................................................................................................... 287 Controls and power supply for furnace. Note the high voltage power supply is activated with the red button, and indicated with a red light. Power for the transducers is provided by a 110 VAC circuit, turned on with a toggle switch and indicated with a green light. Figure B.6. ..................................................................................................................... 288 Blue hoses for gas in and gas out (front bottom) and for coolant water in and out (behind valves) should be inspected for damage such as abrasion or holes. Figure B.7. ..................................................................................................................... 288 The bubbler (arrow), with the hose extending out, should be filled to cover the end of the hose with liquid to prevent air from entering the hose. Figure B.8. ..................................................................................................................... 289 Debris can be seen on the screen in the feed-through leading to the vacuum pump. xxvii Figure B.9. ..................................................................................................................... 290 The three transducers are mounted to copper blocks on the chiller plates underneath the furnace. The mount swivels to allow the angle of the transducer to be adjusted. Figure B.10. ................................................................................................................... 291 Molybdenum sheets to be used as baffles. Figure B.11. ................................................................................................................... 291 Molybdenum baffle being inserted around buffer rods. Figure B.12. ................................................................................................................... 292 Front SALI board of inner furnace. Figure B.13. ................................................................................................................... 292 Front SALI board of inner furnace, inserted into position. Figure B.14. ................................................................................................................... 292 Front refractory bricks of outer furnace. Figure B.15. ................................................................................................................... 292 Front refractory bricks of outer furnace, inserted into position. Figure B.16. ................................................................................................................... 293 Hoist and controls to operate up and down. Note the hoist is mounted to a rail for movement left and right, and that the chain has some slack when the bell jar is in position. Figure B.17. ................................................................................................................... 294 Clamps for the bell jar. The same clamps are used to secure the top and base of the bell jar. Figure B.18. ................................................................................................................... 294 High vacuum valves below the furnace. The gas in and gas out valves should be closed before pumping down the bell jar chamber. Figure B.19. ................................................................................................................... 295 Flow control panel, with flow controls and bypass for the inert gas in and out of the bell jar and the coolant water from the chiller to the base of the transducers. Figure B.20. ................................................................................................................... 296 From left, the furnace over temperature limit controller, furnace temperature controller, and the transducer over temperature limit controller. xxviii Figure B.21. ................................................................................................................... 297 Furnace temperature controller, from left, in run mode, selecting set point 1, and setting set point 1 to 60°C. Figure B.22. ................................................................................................................... 297 208 VAC power switch with red indicator light on, and 110 VAC power switch with green indicator on. Figure B.23. ................................................................................................................... 298 Ammeter and voltmeter for heaters. Resistance decreases with temperature, and the current should be monitored to not exceed 20 A. Operate heater elements at less than 60 V for new elements, less than 120 V for elements that have aged due to use. Figure B.24. ................................................................................................................... 298 Heater ammeter, heater voltmeter, and variac output knob on control box. Figure B.25. ................................................................................................................... 299 Bell jar at rest on side table and furnace opened. There should be some slack in the chain, allowing all the weight of the bell jar to rest on the table. Figure B.26. ................................................................................................................... 301 RUS transducer mount components. Figure B.27. ................................................................................................................... 302 The spiral cut hot zone of the heating elements should be completely contained within the furnace. The power connections to the heating elements should be outside the furnace. Figure B.28. ................................................................................................................... 303 Terminal strip on back of furnace for R-type thermocouples, with thermocouple wires extending into the furnace. Figure C.1. ..................................................................................................................... 303 Gantry and full assembly. Figure C.2. ..................................................................................................................... 304 Gantry and full assembly BOM. Figure C.3. ..................................................................................................................... 305 Gantry. Figure C.4. ..................................................................................................................... 306 Gantry BOM. xxix Figure C.5. ..................................................................................................................... 307 Full assembly. Figure C.6. ..................................................................................................................... 308 Full assembly BOM. Figure C.7. ..................................................................................................................... 309 Bar clamp. Figure C.8. ..................................................................................................................... 310 Bar clamp. Figure C.9. ..................................................................................................................... 311 Bell jar base. Figure C.10. ................................................................................................................... 312 Bell jar base assembly. Figure C.11. ................................................................................................................... 313 Bell jar base assembly BOM. Figure C.12. ................................................................................................................... 314 Bell jar base weldment. Figure C.13. ................................................................................................................... 315 Bell jar base weldment BOM. Figure C.14. ................................................................................................................... 316 Bell jar top. Figure C.15. ................................................................................................................... 317 Cold plate mount 1. Figure C.16. ................................................................................................................... 318 Cold plate mount 2. Figure C.17. ................................................................................................................... 319 Flange tube. Figure C.18. ................................................................................................................... 320 Furnace box. Figure C.19. ................................................................................................................... 321 Furnace box BOM. xxx Figure C.20. ................................................................................................................... 322 Furnace support plate. Figure C.21. ................................................................................................................... 323 Heater support 2. Figure C.22. ................................................................................................................... 324 Heater support. Figure C.23. ................................................................................................................... 325 Inner furnace back SALI board. Figure C.24. ................................................................................................................... 326 Inner furnace base SALI board. Figure C.25. ................................................................................................................... 327 Inner furnace front SALI board. Figure C.26. ................................................................................................................... 328 Inner furnace side SALI board. Figure C.27. ................................................................................................................... 329 Inner furnace top SALI board. Figure C.28. ................................................................................................................... 330 Inner furnace. Figure C.29. ................................................................................................................... 331 Inner furnace BOM. Figure C.30. ................................................................................................................... 332 Outer furnace base center. Figure C.31. ................................................................................................................... 333 Outer furnace base left. Figure C.32. ................................................................................................................... 334 Outer furnace base right. Figure C.33. ................................................................................................................... 335 Outer furnace base. Figure C.34. ................................................................................................................... 336 Outer furnace base BOM. xxxi Figure C.35. ................................................................................................................... 337 Outer furnace front 2. Figure C.36. ................................................................................................................... 338 Outer furnace front. Figure C.37. ................................................................................................................... 339 Outer furnace side 2. Figure C.38. ................................................................................................................... 340 Outer furnace side. Figure C.39. ................................................................................................................... 341 Outer furnace top center. Figure C.40. ................................................................................................................... 342 Outer furnace top left. Figure C.41. ................................................................................................................... 343 Outer furnace top right. Figure C.42. ................................................................................................................... 344 Outer furnace top. Figure C.43. ................................................................................................................... 345 Outer furnace top BOM. Figure C.44. ................................................................................................................... 346 Bell jar top. Figure C.45. ................................................................................................................... 347 RUS and furnace. Figure C.46. ................................................................................................................... 348 RUS clamp. Figure C.47. ................................................................................................................... 349 RUS mount assembly. Figure C.48. ................................................................................................................... 350 RUS mount assembly BOM. Figure C.49. ................................................................................................................... 351 RUS mount. xxxii Figure C.50. ................................................................................................................... 352 RUS support leg. Figure C.51. ................................................................................................................... 353 RUS support plate. Figure C.52. ................................................................................................................... 354 RUS support. Figure C.53. ................................................................................................................... 355 Splash shield mount. Figure C.54. ................................................................................................................... 356 Splash shield. Figure C.55. ................................................................................................................... 357 Standoff. Figure C.56. ................................................................................................................... 358 Table top. Figure C.57. ................................................................................................................... 359 Transducer cold cap. Figure C.58. ................................................................................................................... 360 Valve assembly. Figure C.59. ................................................................................................................... 361 Valve mount. xxxiii 1 Introduction Bulk thermoelectric (TE) materials have been developed with increasing figure of merit, ZT, with some near or exceeding ZT of 2 [Biswas 2012; Heremans 2008], and with improvements in the ZT made by the incorporation of nanoparticles [Androulakis 2007; Zhao 2012a; Zhou 2008a; Alleno 2009; Ji 2007] or nanostructured bulk materials [Toprak 2004; Mi 2008]. While the mechanical properties of many of the undoped TE materials without nanoparticles or nanostructures have been measured and published, doping [Gelbstein 2008], alloying [Gelbstein 2008; Darrow 1969], and incorporation of nanoparticles [Zhao 2008; Kvetková 2013] or nanoprecipitates [Ni 2010; Ren 2010] have each been shown to change the mechanical properties such as hardness or elastic moduli. In addition, the hardness and elastic moduli of TE materials are a function of porosity [Ni 2009]. Furthermore, porosity may be structured similar to particles and precipitates to engineer the properties of the material, such as has been done to improve fatigue life in thermal barrier coatings [Giolli 2009], to improve the operational life of a TE device. As the TE materials are optimized to achieve a high figure of merit, the mechanical properties of the materials also need to be tested. Furthermore, the mechanical properties may be optimized without significant change to the ZT such as by additions of silicon carbide nanoparticles, or in some cases, improving the ZT by the addition of nanoparticles [Zhao 2008]. Thus, the mechanical properties of TE materials must be tested and optimized together with the ZT. 1.1 Applications of thermoelectric materials Thermoelectric materials are capable of being used for either power generation from a temperature gradient or for producing a temperature gradient by applying an 1 electrical potential. TE materials such those based on Mg2Si [Gao 2014], SnTe [Vedeneev 1998; Gelbstein 2009; Leontyev 2012; Gojaev 2010], PbS [Zhao 2012b], PbSe [Zhao 2013], and CoSb3-based skutterudites [Chubilleau 2012; Salvador 2009; Yang 2007; Nolas 1999] are considered candidate materials for waste heat recovery, such as from automotive exhaust heat. Other TE materials such as YbAl3-based compounds are being developed as possible cryogenic Peltier coolers [Lehr 2013a; Lehr 2013b]. TE materials for waste heat recovery applications are subject to stresses from multiple sources when placed in service. For waste heat recovery, the hot side of a TE device would be placed in thermal contact with a source of heat, from sources such as automotive or other engine exhaust, industrial boilers, or metal refineries. The waste heat source is typically not at constant temperature, subjecting the device to stresses caused by thermal gradients, thermal expansion mismatch, thermal transients, in addition to externally applied stresses. In order to understand the mechanical response to these stresses, the mechanical properties of fracture toughness, hardness and elastic moduli must be understood. In addition, the material must be examined for reactions such as decomposition or bloating for the possible operating temperature range. 1.2 Thermoelectric efficiency Thermoelectric (TE) materials are typically brittle materials, heavily doped semiconductor devices or semimetals, which may be used in either power generation or heating or cooling. Typically, the efficiency of a TE material is characterized by the dimensionless figure of merit, ZT, defined as [Tritt 2011; Bux 2010] (1.1) 2 where S is Seebeck coefficient, σ is the electrical conductivity, T is temperature, and κ is thermal conductivity. In heat recovery, the efficiency of a TE device, η, is defined as [Tritt 2011; Bux 2010] ( √ √ ) (1.2) where TH is the hot side temperature and TC is the cold side temperature. The η for the TE material is proportional to the Carnot efficiency, ηC [Tritt 2011]. (1.3) From equations (1.2) and (1.3), it is seen that in order to achieve the efficiency closest to Carnot in a TE device, the ZT must be maximized. Improvements in ZT have incorporated one of several processes, including atomic scale doping or substitution [Gelbstein 2008; Zhu 2009; Zhao 2012b], creation of nanostructured bulk materials [Toprak 2004; Mi 2008; Hsu 2004; Zhu 2009], incorporation of nanoparticles [Androulakis 2007; Zhao 2012a; Zhou 2008a; Alleno 2009; Ji 2007], or a combination of atomic, nanostructural, and microstructural length scale modifications [Biswas 2012]. Until the mid-1990’s, the practical limit for ZT was 1 [Tritt 2011], although recent advancements have demonstrated bulk TE materials with ZT > 2, such as PbTe-based compounds [Biswas 2012] and SnSe [Zhao 2014], along with several other TE materials with ZT > 1, such as skutterudites with ZT of 1.7 [Shi 2011], PbSe with ZT of 1.6 [Zhao 2013], and PbS with ZT of 1.3[Zhao 2012b]. 1.3 Mechanical properties of thermoelectric materials TE materials are typically incorporated into thermoelectric modules, where a series of TE materials in the form of legs (Figure 1.1) are electrically connected in series, 3 but each leg is thermally in parallel [Bux 2010]. Thus, each leg of a TE module is subject to the same temperature gradient individually, but the failure of a single TE leg will result in the failure of an entire module of 10’s or 100’s of legs. Therefore, understanding the mechanical stresses that develop due to thermal gradients, thermal transients, and applied external loads, and how the material will respond to the stresses are critically important to a reliable TE device. The ZT of a TE material is a function of temperature, and therefore different materials achieve maximum efficiency at different temperature ranges. Maximizing efficiency depends on the material choice as well as the operational temperature. An improvement may also be made by segmenting a TE leg, using different TE materials within a leg based on the expected temperature at a section of a leg [Crane 2009]. For example, replacing a part of a Bi2Te3 leg with either TAGS [(AgSbTe2)1-x(GeTe)x] or PbTe on the high temperature side of the leg, depending on whether the leg was p-type or n-type doped, was used to fabricate a TE generator with η > 0.1 for a temperature difference of 500°C [Crane 2009]. A segmented design requires additional modeling to ensure mechanical reliability because of the added complexity of two TE materials in a single leg [Crane 2009]. Numerical simulation of the mechanical response of TE materials requires knowledge of the mechanical properties, such as elastic moduli, hardness or fracture toughness of the material [Kardestuncer 1987]. These mechanical properties are typically not measured for any case of optimized TE materials. The only published elastic moduli for TE materials are typically for undoped or single crystal case, such as elastic moduli in SnTe [Beattie 1969] or PbS [Dalven 1969; Bhagavantam 1951], which may not be 4 comparable to doped or otherwise optimized TE materials. Thus, to understand the mechanical response and to run simulation results, the mechanical properties, including elastic moduli, hardness, and fracture toughness, must be measured on the optimized TE material. The calculation of thermal stresses in a temperature transient may be demonstrated with the simplified case of a flat plate. The maximum surface stress on a flat plate, Smax, is a function of the temperature-dependent Young’s modulus, E(T), Poisson’s ratio, (T), thermal expansion, (T), and a function of the Biot modulus, f(Bi), where [Case 2012] Smax  E T  αT   T  f(Bi) 1   T  (1.4) The function f(Bi) increases monotonically with increasing Biot modulus Bi, where Bi = ah/κ, and a is the characteristic dimension of the specimen, h is the surface heat transfer coefficient and κ is the thermal conductivity of the specimen. In this case, the elastic moduli, E(T) and (T), and thermal expansion, (T), are required mechanical/thermal properties to determine the stress in the material. 5 Figure 1.1. Schematic of a thermoelectric couple, indicating current flow when placed in a temperature gradient. 6 1.3.1 Elastic moduli The elastic moduli of a specimen are required to determine the elastic response of a material to a given stress. The moduli may change as a material system is optimized as a TE material, such as by the addition of particles or precipitates, or through dopants. Additions of dopants and elements to produce solid-state precipitates have been commonly applied to thermoelectric materials. Lead telluride-based thermoelectrics have achieved record efficiencies in part by using this technique [Zhao 2013]. Also, if microcrack damage is accumulated, through thermal cycling or mechanical loading, the microcrack-induced reduction in E may be used to monitor the extent of the microcrack damage, as has been previously shown for non-TE materials [Fan 2012; Case 1993; Ghassemi Kakroudi 2009]. For the elastic moduli, additions of well-connected, insoluble nanoparticles may be modeled several ways, including rule of mixtures [Hashin 1962], constant strain [Hashin 1962], Hashin particulate composite model [Bedolla 2012; Couturier 1997; Hashin 1962], and Halpin-Tsai composite [Halpin 1992]. Each of these models ignores the possible effects of the size of nanoparticles within a matrix, but only the rule of mixtures model predicts a significant change in elastic modulus for small additions of nanoparticles. Additions of silver nanoparticles (AgNP) have improved the ZT of barium-filled skutterudite by 30% [Zhou 2012], which raised interest in examining the mechanical properties of the material. The Young’s modulus of silver and CoSb3-based skutterudite are not dramatically different (E = 82 GPa for Ag [Neighbours 1958; Chang 1966], E = 109 GPa to 148 GPa for SKD [Rogl 2011]). Therefore, addition of 0.5 wt% silver to skutterudite should not change the E significantly, if the models for additions insoluble 7 nanoparticles are accurate [Hashin 1962; Bedolla 2012; Couturier 1997; Halpin 1992]. However, addition of 0.5 wt% silver to skutterudite was accompanied by a modest increase in Seebeck along with the expected change in electrical conductivity [Zhou 2012]. The change in Seebeck could indicate the nanoparticle additions are not be completely insoluble, and therefore there may be a change in the mechanical properties of the TE material as well that the models would not predict. The addition of silicon carbide nanoparticles (SiCNP) or precipitates to a TE material may improve various material properties, such as acting as a scattering site to reduce the thermal conductivity [Zhou 2008b; Li 2006], or to inhibit bloating and creep behavior [Ni 2013]. The ZT of Bi2Te3 was improved from 0.99 to 1.04, while the hardness, fracture toughness and E were each increased by the addition of 0.1 vol% to 1.0 vol% SiCNP [Zhao 2008], demonstrating the possibilities of SiCNP or other particle additions to TE materials. Dopants, as well as solid solution changes or modifications to stiochiometry, may have a strong influence on the elastic moduli. In a review by Ren et al., ratios between the maximum and minimum moduli, Emax/Emin and Bmax/Bmin, of up to 3 were observed for changes in solid solution [Ren 2007]. If additions such as AgNP or SiCNP form a solid solution with a TE material, or if the composition of the TE material is otherwise altered, the mechanical properties of the TE material could have significant changes. 1.3.2 Fracture toughness In order to create a reliable TE device, the TE material must withstand the environmental stresses and the cycling of the intended application without failure. In general, TE materials are brittle, including Zn4Sb3 [Ueno 2005], CoSb3 [Salvador 2009, 8 Yang 2009a], PbTe [Salvador 2009], and lead antimony silver tellurium compounds (LAST) [Ren 2006]. This generally implies that the mechanical failure mode of TE materials is fracture, similar to ceramics [Kingery 1976]. The conditions under which crack propagation may occur from a flaw of size a [Rice 1998] are (6) √ where σfract is the fracture stress, KC is the fracture toughness of a material, and Y is a shape factor based upon the geometry of the flaw. Therefore, KC is a necessary material property a designer requires to minimize the risk of fracture in a device. Unfortunately, the KC of TE materials is typically low, between 0.5 and 1.5 MPam1/2 or less [Eilertsen 2013; Fan 2012; Ni 2013]. However, fracture toughness may be changed by the addition of particles or second phases. The fracture toughness of Bi2Te3 was increased from 1.14 MPa-m1/2 to 1.35 MPa-m1/2 by the addition of 0.1 vol% SiCNP [Zhao 2008]. Furthermore, additions of nanoparticles may influence fracture mode or sintering behavior [Mukhopadhyay 2010; Ni 2013], and a fracture mode change has been accompanied by either an increase [Mukhopadhyay 2010; Kawabata 1977; Karakasidis 2011] or decrease [Yamada 2010; Jang 2010] in KC. 1.4 Porosity effects on mechanical properties The addition of porosity to a specimen may have profound effects. The elastic modulus of several TE materials drops off according to an empirically derived exponential relationship, (1.4a) 9 where E is Young’s modulus, E0 is the modulus of theoretically dense specimen, P is volume fraction porosity, and b is a fitting parameter, commonly observed to be between 2 and 6 for a wide range of solid materials [Rice 1998]. This exponential relationship may be extended to include the shear modulus, G, and bulk modulus, B, of TE materials [Ni 2009], as well as hardness, H [Ni 2011; Ren 2008], as the general equation (1.4b) where A represents the property E, G, B, or H [Rice 1998]. If the term bP is small, the equation may be linearized by using the first two terms of the Taylor series expansion of the exponential to give (1.4c) For the materials in this study, this linear approximation is appropriate for the data. The fracture toughness of a brittle material may increase, decrease or remain constant with increasing porosity [Rice 1995]. In particular, an evaluation including Al2O3, B4C and Si3N4 at P = 0.1 to P = 0.15 indicated KIC values as high or higher than at P = 0 [Rice 1995]. The maintenance of KIC values with increasing P may be due to the pore blunting or deflecting the crack tip [Case 2012]. 1.5 Temperature effects on mechanical properties The elastic moduli decreases with increasing temperature. For temperatures greater than 0.3 to 0.5 of the Debye temperature, the temperature-dependent change may be represented by [Schmidt 2012] – – (1.5a) – – (1.5b) 10 where ERT and GRT are the room temperature Young’s and shear moduli respectively, and the constants bTE and bTG are experimentally measured fitting factors. A change in the slope, bTE and bTG, indicates a change within the material, such as the onset of grain boundary sliding or an order-disorder transition [Schmidt 2012; Ren 2009; Wachtman 2009]. These changes within a material may be significant, and in some cases may be controlled by particle additions to the material [Schmidt 2012]. 1.6 Sintering of thermoelectric materials Typically, thermoelectric materials are fabricated by sintering of a powder processed ingot, either by hot pressing (HP) or pulsed electric current sintering (PECS) [Ni 2011]. These two methods of sintering both may produce dense specimens, but the microstructure, including pore shape and distribution, grain size and grain growth, may vary considerably [Ni 2011]. The HP method is well established, but the PECS method may permit consolidation in considerably shorter time, reducing cost of sintering and reducing time for grain growth [Savary 2012; Recknagel 2007; Zhao 2012b; Biswas 2012]. Additionally, the use of sintering aids to create a small, < 1%, amount of liquid phase at sintering temperature may permit the use of lower sintering temperature and shorter time to densify a specimen [Barsoum 2003]. For TE materials, the mechanical properties of elastic moduli, hardness and fracture toughness for specimens sintered by each of HP and PECS have not previously been explored. Furthermore, additions of sintering aids have not been explored for Sbbased skutterudite TE materials, nor for most TE materials. The exploring of sintering aids and creating a process for the development of new sintering aids may improve the sintering time and cost [Barsoum 2003]. 11 REFERENCES 12 REFERENCES [Alleno, Chen, Chubilleau, Lenoir, Rouleau, Trichet, Villeroy 2009] Alleno, E., Chen, L., Chubilleau, C., Lenoir, B., Rouleau, O., Trichet, M.F., Villeroy, B. Thermal Conductivity Reduction in CoSb3–CeO2 Nanocomposites. Journal of Electronic Materials 39 (2009) 1966–1970. [Androulakis, Lin, Kong, Uher, Wu, Hogan, Cook, Caillat, Paraskevopoulos, Kanatzidis 2007] Androulakis, J., Lin, C.-H., Kong, H.-J., Uher, C., Wu, C.-I., Hogan, T., Cook, B.A., Caillat, T., Paraskevopoulos, K.M., Kanatzidis, M.G. Spinodal decomposition and nucleation and growth as a means to bulk nanostructured thermoelectrics: enhanced performance in Pb(1-x)Sn(x)Te-PbS. Journal of the American Chemical Society 129 (2007) 9780–8. [Barsoum 2003] Barsoum, M.W. Fundamentals of Ceramics. Taylor & Francis Group, New York, 2003. [Beattie 1969] Beattie, A.G. Temperature Dependence of the Elastic Constants of Tin Telluride. Journal of Applied Physics 40 (1969) 4818. [Bedolla, Lemus-Ruiz, Contreras 2012] Bedolla, E., Lemus-Ruiz, J., Contreras, A. Synthesis and characterization of Mg-AZ91/AlN composites. Materials and Design 38 (2012) 91–98. [Bhagavantam, Rao 1951] Bhagavantam, S., Rao, T. Elastic constants of galena. Nature 168 (1951) 42. [Biswas, He, Blum, Wu, Hogan, Seidman, Dravid, Kanatzidis 2012] Biswas, K., He, J., Blum, I.D., Wu, C.-I., Hogan, T.P., Seidman, D.N., Dravid, V.P., Kanatzidis, M.G. High-performance bulk thermoelectrics with all-scale hierarchical architectures. Nature 489 (2012) 414–8. [Bux, Fleurial, Kaner 2010] Bux, S.K., Fleurial, J.-P., Kaner, R.B. Nanostructured materials for thermoelectric applications. Chemical communications (Cambridge, England) 46 (2010) 8311–24. [Case 2012] Case, E.D. Thermo-mechanical properties of thermoelectric materials. In: D.M. Rowe, ed., Thermoelectrics and its Energy Harvesting: Modules, Systems and Applications. CRC Press, Boca Raton, Florida, 2012, 581. [Case, Kim 1993] Case, E.D., Kim, Y. The effect of surface-limited microcracks on the effective Young’s modulus of ceramics. Journal of Materials Science 28 (1993) 1885–1900. 13 [Chang, Himmel 1966] Chang, Y.A., Himmel, L. Temperature Dependence of the Elastic Constants of Cu, Ag, and Au above Room Temperature. Journal of Applied Physics 37 (1966) 3567–3572. [Chubilleau, Lenoir, Masschelein, Dauscher, Candolfi, Guilmeau, Godart 2012] Chubilleau, C., Lenoir, B., Masschelein, P., Dauscher, a., Candolfi, C., Guilmeau, E., Godart, C. High temperature thermoelectric properties of CoSb3 skutterudites with PbTe inclusions. Journal of Materials Science 48 (2012) 2761–2766. [Couturier, Ducret, Merle, Disson, Joubert 1997] Couturier, R., Ducret, D., Merle, P., Disson, J.P., Joubert, P. Elaboration and characterization of a metal matrix composite: Al/AlN. Journal of the European Ceramic Society 17 (1997) 1861–1866. [Crane, Kossakovski, Bell 2009] Crane, D.T., Kossakovski, D., Bell, L.E. Modeling the Building Blocks of a 10% Efficient Segmented Thermoelectric Power Generator. Journal of Electronic Materials 38 (2009) 1382–1386. [Dalven 1969] Dalven, R. A review of the semiconductor properties of PbTe, PbSe, PbS and PbO. Infrared physics 9 (1969) 141–184. [Darrow, White, Roy 1969] Darrow, M., White, W., Roy, R. Micro-indentation hardness variation as a function of composition for polycrystalline solutions in the systems PbS/PbTe, PbSe/PbTe, and PbS/PbSe. Journal of Materials Science 4 (1969) 313– 319. [Eilertsen, Subramanian, Kruzic 2013] Eilertsen, J., Subramanian, M. a., Kruzic, J.J. Fracture toughness of Co4Sb12 and In0.1Co4Sb12 thermoelectric skutterudites evaluated by three methods. Journal of Alloys and Compounds 552 (2013) 492–498. [Fan, Case, Ren, Shu, Baumann 2012] Fan, X., Case, E.D., Ren, F., Shu, Y., Baumann, M.J. Part II: fracture strength and elastic modulus as a function of porosity for hydroxyapatite and other brittle materials. Journal of the mechanical behavior of biomedical materials 8 (2012) 99–110. [Gao, Berkun, Schmidt, Luzenski, Lu, Bordon Sarac, Case, Hogan 2014] Gao, P., Berkun, I., Schmidt, R.D., Luzenski, M.F., Lu, X., Bordon Sarac, P., Case, E.D., Hogan, T.P. Transport and Mechanical Properties of High-ZT Mg2.08Si0.4−x Sn0.6Sb x Thermoelectric Materials. Journal of Electronic Materials 43 (2014) 1790–1803. [Gelbstein 2009] Gelbstein, Y. Thermoelectric power and structural properties in twophase Sn/SnTe alloys. Journal of Applied Physics 105 (2009) 023713. 14 [Gelbstein, Gotesman, Lishzinker, Dashevsky, Dariel 2008] Gelbstein, Y., Gotesman, G., Lishzinker, Y., Dashevsky, Z., Dariel, M.P. Mechanical properties of PbTe-based thermoelectric semiconductors. Scripta Materialia 58 (2008) 251–254. [Ghassemi Kakroudi, Yeugo-Fogaing, Huger, Gault, Chotard 2009] Ghassemi Kakroudi, M., Yeugo-Fogaing, E., Huger, M., Gault, C., Chotard, T. Influence of the thermal history on the mechanical properties of two alumina based castables. Journal of the European Ceramic Society 29 (2009) 3197–3204. [Giolli, Scrivani, Rizzi, Borgioli, Bolelli, Lusvarghi 2009] Giolli, C., Scrivani, a., Rizzi, G., Borgioli, F., Bolelli, G., Lusvarghi, L. Failure Mechanism for Thermal Fatigue of Thermal Barrier Coating Systems. Journal of Thermal Spray Technology 18 (2009) 223–230. [Gojaev, Murguzov, Mamedov, Abdurragimov 2010] Gojaev, E.M., Murguzov, M.I., Mamedov, I.M., Abdurragimov, a. a. Synthesis and properties of SnTe-PrTe solid solutions. Inorganic Materials 46 (2010) 1061–1064. [Halpin 1992] Halpin, J.C. Primer on composite materials analysis. Technomic Publishing Company, Inc., Lancaster, Pennsylvania, 1992. [Hashin 1962] Hashin, Z. The Elastic Moduli of Heterogeneous Materials. Journal of Applied Mechanics 29 (1962) 143–150. [Heremans, Jovovic, Toberer, Saramat, Kurosaki, Charoenphakdee, Yamanaka, Snyder 2008] Heremans, J.P., Jovovic, V., Toberer, E.S., Saramat, A., Kurosaki, K., Charoenphakdee, A., Yamanaka, S., Snyder, G.J. Enhancement of thermoelectric efficiency in PbTe by distortion of the electronic density of states. Science (New York, N.Y.) 321 (2008) 554–7. [Hsu, Loo, Guo, Chen, Dyck, Uher, Hogan, Polychroniadis, Kanatzidis 2004] Hsu, K.F., Loo, S., Guo, F., Chen, W., Dyck, J.S., Uher, C., Hogan, T., Polychroniadis, E.K., Kanatzidis, M.G. Cubic AgPb(m)SbTe(2+m): bulk thermoelectric materials with high figure of merit. Science (New York, N.Y.) 303 (2004) 818–21. [Jang, Kim, Han, Seo, Hong, Woo, Sakka 2010] Jang, B.-K., Kim, S.-Y., Han, I.-S., Seo, D.-W., Hong, K.-S., Woo, S.-K., Sakka, Y. Influence of uni and bi-modal SiC composition on mechanical properties and microstructure of reaction-bonded SiC ceramics. Journal of the Ceramic Society of Japan 118 (2010) 1028–1031. [Ji, He, Alboni, Su, Gothard, Zhang, Tritt, Kolis 2007] Ji, X., He, J., Alboni, P., Su, Z., Gothard, N., Zhang, B., Tritt, T.M., Kolis, J.W. Thermal conductivity of CoSb3 nano-composites grown via a novel solvothermal nano-plating technique. physica status solidi RRL – Rapid Research Letters 1 (2007) 229–231. 15 [Karakasidis, Charitidis 2011] Karakasidis, T.E., Charitidis, C. a. Influence of nanoinclusions’ grain boundaries on crack propagation modes in materials. Materials Science and Engineering: B 176 (2011) 490–493. [Kardestuncer, Norrie 1987] Finite element handbook. McGraw-Hill, New York, 1987. [Kawabata, Izumi 1977] Kawabata, T., Izumi, O. The relationship between fracture toughness and transgranular fracture in an Al-6.0% Zn-2.5% Mg alloy. Acta Metallurgica 25 (1977) 505–512. [Kingery, Bowen, Uhlman 1976] Kingery, W.D., Bowen, H.K., Uhlman, D.R. Introduction to Ceramics. John Wiley & Sons, New York, 1976. [Kvetková, Duszová, Kašiarová, Dorčáková, Dusza, Balázsi 2013] Kvetková, L., Duszová, A., Kašiarová, M., Dorčáková, F., Dusza, J., Balázsi, C. Influence of processing on fracture toughness of Si3N4+graphene platelet composites. Journal of the European Ceramic Society 33 (2013) 2299–2304. [Lehr, Morelli 2013a] Lehr, G.J., Morelli, D.T. Thermoelectric Properties of Yb1−x (Er,Lu) x Al3 Solid Solutions. Journal of Electronic Materials 42 (2013) 1697– 1701. [Lehr, Morelli, Jin, Heremans 2013b] Lehr, G.J., Morelli, D.T., Jin, H., Heremans, J.P. Enhanced thermoelectric power factor in Yb1−xScxAl2 alloys using chemical pressure tuning of the Yb valence. Journal of Applied Physics 114 (2013) 223712. [Leontyev, Ivanova, Bente, Gremenok 2012] Leontyev, V.G., Ivanova, L.D., Bente, K., Gremenok, V.F. Fabrication of bulk materials from fine particles of PbTe-SnTe solid solutions prepared through thermal decomposition of salts. Inorganic Materials 48 (2012) 991–996. [Li, Liu 2006] Li, J.-F., Liu, J. Effect of nano-SiC dispersion on thermoelectric properties of Bi 2 Te 3 polycrystals. Physica Status Solidi (a) 203 (2006) 3768–3773. [Mi, Zhao, Zhu, Tu 2008] Mi, J.L., Zhao, X.B., Zhu, T.J., Tu, J.P. Thermoelectric properties of Yb 0.15 Co 4 Sb 12 based nanocomposites with CoSb 3 nanoinclusion. Journal of Physics D: Applied Physics 41 (2008) 205403. [Mukhopadhyay, Chakravarty, Basu 2010] Mukhopadhyay, A., Chakravarty, D., Basu, B. Spark Plasma-Sintered WC-ZrO 2 -Co Nanocomposites with High Fracture Toughness and Strength. Journal of the American Ceramic Society 1763 (2010) 1754–1763. [Neighbours, Alers 1958] Neighbours, J., Alers, G. Elastic constants of silver and gold. Physical Review 885 (1958) 707–712. 16 [Ni, Case, Khabir, Stewart, Wu, Hogan, Timm, Girard, Kanatzidis 2010] Ni, J.E., Case, E.D., Khabir, K.N., Stewart, R.C., Wu, C.-I., Hogan, T.P., Timm, E.J., Girard, S.N., Kanatzidis, M.G. Room temperature Young’s modulus, shear modulus, Poisson's ratio and hardness of PbTe–PbS thermoelectric materials. Materials Science and Engineering: B 170 (2010) 58–66. [Ni, Case, Schmidt, Wu, Hogan, Trejo, Lara-Curzio, Kanatzidis 2013] Ni, J.E., Case, E.D., Schmidt, R.D., Wu, C.-I., Hogan, T.P., Trejo, R.M., Lara-Curzio, E., Kanatzidis, M.G. Fracture mode, microstructure and temperature-dependent elastic moduli for thermoelectric composites of PbTe–PbS with SiC nanoparticle additions. Philosophical Magazine 93 (2013) 4412–4439. [Ni, Case, Stewart, Wu, Hogan, Kanatzidis 2011] Ni, J.E., Case, E.D., Stewart, R., Wu, C.-I., Hogan, T.P., Kanatzidis, M.G. Bloating in (Pb0.95Sn0.05Te)0.92(PbS)0.080.055%PbI2 Thermoelectric Specimens as a Result of Processing Conditions. Journal of Electronic Materials 41 (2011) 1153–1158. [Ni, Ren, Case, Timm 2009] Ni, J.E., Ren, F., Case, E.E.D., Timm, E.J.E. Porosity dependence of elastic moduli in LAST (lead–antimony–silver–tellurium) thermoelectric materials. Materials Chemistry and Physics 118 (2009) 459–466. [Nolas, Morelli, Tritt 1999] Nolas, G.S., Morelli, D.T., Tritt, T.M. SKUTTERUDITES: A Phonon-Glass-Electron Crystal Approach to Advanced Thermoelectric Energy Conversion Applications. Annual Review of Materials Science 29 (1999) 89–116. [Recknagel, Reinfried, Höhn, Schnelle, Rosner, Grin, Leithe-Jasper 2007] Recknagel, C., Reinfried, N., Höhn, P., Schnelle, W., Rosner, H., Grin, Y., Leithe-Jasper, a. Application of spark plasma sintering to the fabrication of binary and ternary skutterudites. Science and Technology of Advanced Materials 8 (2007) 357–363. [Ren, Case, Ni, Timm 2009] Ren, F., Case, E.D., Ni, J.E., Timm, E.J. Temperaturedependent elastic moduli of lead telluride-based thermoelectric materials. (2009) 37–41. [Ren, Case, Timm, Lara-Curzio, Trejo 2010] Ren, F., Case, E.D., Timm, E.J., LaraCurzio, E., Trejo, R.M. Anomalous temperature-dependent Young’s modulus of a cast LAST (Pb–Sb–Ag–Te) thermoelectric material. Acta Materialia 58 (2010) 31– 38. [Ren, Case, Timm, Schock 2007] Ren, F., Case, E.D., Timm, E.J., Schock, H.J. Young’s modulus as a function of composition for an n-type lead–antimony–silver–telluride (LAST) thermoelectric material. Philosophical Magazine 87 (2007) 4907–4934. 17 [Ren, Case, Timm, Schock 2008] Ren, F., Case, E.D., Timm, E.J., Schock, H.J. Hardness as a function of composition for n-type LAST thermoelectric material. Journal of Alloys and Compounds 455 (2008) 340–345. [Rice 1995] Rice, R.W. Comment on “Effective elastic moduli of porous ceramic materials.”Journal of the American Ceramic Society 78 (1995) 1711. [Rice 1998] Rice, R.W. Porosity of Ceramics. Marcel Dekker, New York, 1998. [Rogl, Rogl 2011] Rogl, G., Rogl, P. Mechanical Properties of Skutterudites. Science of Advanced Materials 3 (2011) 517–538. [Salvador, Yang, Shi, Wang, Wereszczak, Kong, Uher 2009] Salvador, J.R., Yang, J., Shi, X., Wang, H., Wereszczak, a. a., Kong, H., Uher, C. Transport and mechanical properties of Yb-filled skutterudites. Philosophical Magazine 89 (2009) 1517–1534. [Savary, Gascoin, Marinel, Heuguet 2012] Savary, E., Gascoin, F., Marinel, S., Heuguet, R. Spark plasma sintering of fine Mg2Si particles. Powder Technology 228 (2012) 295–300. [Schmidt, Case, Ni, Sakamoto, Trejo, Lara-Curzio 2012] Schmidt, R.D., Case, E.D., Ni, J.E., Sakamoto, J.S., Trejo, R.M., Lara-Curzio, E. Temperature-dependent Young’s modulus, shear modulus and Poisson's ratio of p -type Ce 0.9 Fe 3.5 Co 0.5 Sb 12 and n -type Co 0.95 Pd 0.05 Te 0.05 Sb 3 skutterudite thermoelectric materials. Philosophical Magazine 92 (2012) 727–759. [Shi, Yang, Salvador, Chi, Cho, Wang, Bai, Yang, Zhang, Chen 2011] Shi, X., Yang, J., Salvador, J.R., Chi, M., Cho, J.Y., Wang, H., Bai, S., Yang, J., Zhang, W., Chen, L. Multiple-filled skutterudites: high thermoelectric figure of merit through separately optimizing electrical and thermal transports. Journal of the American Chemical Society 133 (2011) 7837–46. [Toprak, Stiewe, Platzek, Williams, Bertini, Muller, Gatti, Zhang, Rowe, Muhammed 2004] Toprak, M.S., Stiewe, C., Platzek, D., Williams, S., Bertini, L., Muller, E., Gatti, C., Zhang, Y., Rowe, M., Muhammed, M. The Impact of Nanostructuring on the Thermal Conductivity of Thermoelectric CoSb3. Advanced Functional Materials 14 (2004) 1189–1196. [Tritt 2011] Tritt, T.M. Thermoelectric Phenomena, Materials, and Applications. Annual Review of Materials Research 41 (2011) 433–448. [Vedeneev, Krivoruchko, Sabo 1998] Vedeneev, V.P., Krivoruchko, S.P., Sabo, E.P. Tin telluride based thermoelectrical alloys. Semiconductors 32 (1998) 241–244. 18 [Wachtman, Cannon, Matthewson 2009] Wachtman, J.B., Cannon, W.R., Matthewson, M.J. Mechanical Properties of Ceramics. Wiley-Vch, Hoboken, N.J., 2009. [Yamada, Sekine, Kumazawa, Tanabe 2010] Yamada, M., Sekine, K., Kumazawa, T., Tanabe, Y. Relationship between the cone crack and fracture mode in ceramics under high-velocity-projectile impact. 96 (2010) 903–908. [Yang, Meisner, Pedersen 2007] Yang, J., Meisner, G.P., Pedersen, B.L. Low temperature transport and structural properties of misch-metal-filled skutterudites. (2007) 1–7. [Zhao, Hao, Lo, Wu, Zhou, Lee, Li, Biswas, Hogan, Uher, Wolverton, Dravid, Kanatzidis 2013] Zhao, L.-D., Hao, S., Lo, S.-H., Wu, C.-I., Zhou, X., Lee, Y., Li, H., Biswas, K., Hogan, T.P., Uher, C., Wolverton, C., Dravid, V.P., Kanatzidis, M.G. High thermoelectric performance via hierarchical compositionally alloyed nanostructures. Journal of the American Chemical Society 135 (2013) 7364–70. [Zhao, He, Hao, Wu, Hogan, Wolverton, Dravid, Kanatzidis 2012a] Zhao, L.-D., He, J., Hao, S., Wu, C.-I., Hogan, T.P., Wolverton, C., Dravid, V.P., Kanatzidis, M.G. Raising the thermoelectric performance of p-type PbS with endotaxial nanostructuring and valence-band offset engineering using CdS and ZnS. Journal of the American Chemical Society 134 (2012) 16327–36. [Zhao, He, Wu, Hogan, Zhou, Uher, Dravid, Kanatzidis 2012b] Zhao, L.-D., He, J., Wu, C.-I., Hogan, T.P., Zhou, X., Uher, C., Dravid, V.P., Kanatzidis, M.G. Thermoelectrics with earth abundant elements: high performance p-type PbS nanostructured with SrS and CaS. Journal of the American Chemical Society 134 (2012) 7902–12. [Zhao, Lo, Zhang, Sun, Tan, Uher, Wolverton, Dravid, Kanatzidis 2014] Zhao, L.-D., Lo, S.-H., Zhang, Y., Sun, H., Tan, G., Uher, C., Wolverton, C., Dravid, V.P., Kanatzidis, M.G. Ultralow thermal conductivity and high thermoelectric figure of merit in SnSe crystals. Nature 508 (2014) 373–7. [Zhao, Zhang, Li, Zhou, Liu, Liu 2008] Zhao, L.-D., Zhang, B.-P., Li, J.-F., Zhou, M., Liu, W.-S., Liu, J. Thermoelectric and mechanical properties of nano-SiC-dispersed Bi2Te3 fabricated by mechanical alloying and spark plasma sintering. Journal of Alloys and Compounds 455 (2008) 259–264. [Zhou, Li, Kita 2008a] Zhou, M., Li, J.-F., Kita, T. Nanostructured AgPb(m)SbTe(m+2) system bulk materials with enhanced thermoelectric performance. Journal of the American Chemical Society 130 (2008) 4527–32. 19 [Zhou, Li, Kita 2008b] Zhou, M., Li, J.-F., Kita, T. Nanostructured AgPb(m)SbTe(m+2) system bulk materials with enhanced thermoelectric performance. Journal of the American Chemical Society 130 (2008) 4527–32. [Zhou, Wang, Zhang, Chi, Su, Sakamoto, Uher 2012] Zhou, X., Wang, G., Zhang, L., Chi, H., Su, X., Sakamoto, J., Uher, C. Enhanced thermoelectric properties of Bafilled skutterudites by grain size reduction and Ag nanoparticle inclusion. Journal of Materials Chemistry 22 (2012) 2958. [Zhu, Lee, Lan, Wang, Joshi, Wang, Yang, Vashaee, Guilbert, Pillitteri, Dresselhaus, Chen, Ren 2009] Zhu, G., Lee, H., Lan, Y., Wang, X., Joshi, G., Wang, D., Yang, J., Vashaee, D., Guilbert, H., Pillitteri, a., Dresselhaus, M., Chen, G., Ren, Z. Increased Phonon Scattering by Nanograins and Point Defects in Nanostructured Silicon with a Low Concentration of Germanium. Physical Review Letters 102 (2009) 196803. 20 2 Experimental Procedure 2.1 Material preparation For the YbAl3 (chapter 3), Mg2Si (chapters 4 and 5), Ba0.3Co4Sb12 (chapter 6), SnTe (chapter 7), and PbS and PbSe-based thermoelectric materials (chapter 8) in this dissertation, ingots or lump material were prepared from high purity elemental powders at Michigan State University, Northwestern University, University of Michigan, or by Alfa Aesar. YbAl3 ingots (chapter 3) were arc melted at Michigan State University in an Al flux under argon, with each ingot melted at least 5 times to ensure homogeneity. The excess Al was removed by crushing the ingot and etching the Al with a 5 M solution of NaOH for 24 h. The etched powder was passed through a 100-mesh (149 µm) sieve before powder processing or sintering. The Mg2Si lump material (chapters 4 and 5) was commercially prepared (45518, Alfa Aesar, Ward Hill, MA) from high purity elements (99.99% metals basis). The lump material was crushed, ground, sieved and reground (CGSR), as described in section 2.2. Barium-doped skutterudite ingots, Ba0.3Co4Sb12 (chapter 6), were fabricated from elemental Ba (pieces, 99.9% pure), Co (powder 99.5% pure), and Sb (shot 99.999% pure), by sealing the elements in evacuated carbon coated quartz ampoules, heated to 1373 K for 5 h, quenched, then annealing at 1023 K for 7 days. The skutterudite ingot produced was processed by CGSR, as described in section 2.2. SnTe (chapter 7) ingots were prepared by combining high purity elemental powders of Sn (Plamaterials purity 99.999 %) and Te (5 N Plus Inc., purity 99.999 %) in a quartz ampoule and sealed under vacuum. The sealed tubes were heated to either 1273 21 K for 10 h or 1148 K for 12 h, then quenched. The ingot material produced was processed by CGSR, as described in section 2.2. Na-doped PbS and PbSe ingots (chapter 8) with x% CdS/ZnS, where the at fraction x = 1, 2, 3 and 4, were produced by sealing high purity elements in evacuated carbon coated quartz ampoules and heating to 1473 K for 4 h, then quenched. The resulting ingots were crushed, and the crushed powder was passed through a 53 µm sieve before sintering. 2.2 Crushing, grinding, sieving and regrinding (CGSR) process The Mg2Si lump, Ba0.3Co4Sb12 sintered billets for reprocessing and SnTe ingot (chapters 4 – 7) were crushed, ground, sieved and reground (CGSR), using a process originally developed for another TE material, LAST (lead-antimony-silver-tellurium) [Pilchak 2007], and previously used for Pb0.95Sn0.05Te-PbS 8% [Ni 2010] and other Sbbased skutterudites [Schmidt 2011]. All CGSR processing was performed in a glove box under argon (Omni-Lab double glove box) equipped with an oxygen sensor and a moisture sensor (Vacuum Atmospheres Company, Hawthorne, CA). The lump or ingot material was ground for 5 minutes in a WC-lined mechanical mortar and pestle machine (Retsch RM200, Retsch GmbH, Haan, Germany). The powder was transferred to a sieve, 53 μm ASTM E 11 sieve (60.150.000053, Retsch GmbH, Haan, Germany) for the Mg2Si and SnTe, or 75 μm sieve (60.150.000075, Retsch GmbH) for the Ba0.3Co4Sb12, with a bottom collector pan (697203050, Retsch GmbH). The top of the sieve was covered with a sheet of aluminum foil to limit possible contamination, and then the ground powder was passed through the sieve on a shaker table (AS 200, Retsch GmbH) for 5 to 15 minutes. The sieved powder was collected and 22 stored in glass scintillation vials (RPI 121000, Research Products International Corp., Mt. Prospect, IL), and the powder that did not pass was returned to the mortar for regrinding. Up to 5 grindings were performed per batch of powder to pass the ground lump or ingot through the sieve. Sieves were cleaned between uses with different material systems or when the amount of ground powder noticeably decreased to indicate clogging. The sieve was placed in an ultrasonic cleaner (VWR, 98000-336) with RO water and detergent. A drop of ethanol was used to break the water surface tension and remove any large bubbles trapped under the mesh before cleaning. The ultrasonic cleaner was run for 5 minutes, the screen was rinsed with ethanol and RO water, then dried in a 353 K oven for 20 to 30 minutes. 2.3 Milling of CGSR powder The CGSR powder was milled by one of two methods, planetary ball milling or vibratory milling. 2.3.1 Planetary ball milling Planetary ball milling (PM) was performed on Mg2Si CGSR powder and SnTe CGSR powder in the same glove box under argon (Omni-Lab double glove box) as the CGSR processing of powder was performed, allowing the material to remain in the same argon atmosphere for the entire powder processing process. The CGSR powder was milled in 10–30 g batches with 100 g of 10-mm diameter spherical alumina media at 150 rpm for 3 h (planetary mill PM100 and 250 mL alumina jar 01.462.0221, Retsch GmbH, Haan, Germany). The powder was collected from the mill by transferring the contents of the mill jar to a coarse sieve (60.150.000, Retsch 23 GmbH) stacked on a bottom collector pan (697203050, Retsch GmbH). The jar was scraped down with a metal spatula (82027-532, VWR, Radnor, PA) to collect material caked to the walls of the jar. The top of the sieve was covered with a sheet of aluminum foil, then the loose powder on the media was removed on a shaker table for 5 to 15 minutes. The milled powder was collected and stored in glass scintillation vials. Cleaning of the PM jar and media was performed when changing the TE material milled, or when the jar was removed from the glove box. If the jar was used for Mg2Si, the jar and media were first soaked overnight in reverse osmosis (RO) water, then rinsed with diluted HCl and RO water to reduce the possibility of a significant reaction between the TE material and the cleaning chemicals. The PM jar and media were cleaned using aqua regia (3-4 parts HCl: 1 part HNO3) in a fume hood while wearing appropriate personal protective equipment. The media and jar were soaked in aqua regia until no remaining TE material was visible, and for at least 20 minutes. Stirring and applying of aqua regia was performed with disposable plastic pipets (14670-147, VWR). The used aqua regia was disposed of in a vented container. The media and jar were thoroughly rinsed with RO water, filling the jar multiple times with RO water, before air drying. 2.3.2 Vibratory ball milling Vibratory ball milling (VM) was performed on Mg2Si CGSR powder (chapter 5), Ba0.3Co4Sb12 CGSR powder (chapter 6), and YbAl3 sieved powder (chapter 3). For the Mg2Si CGSR powder and Ba0.3Co4Sb12 CGSR powder, each batch of powder (5 g for Mg2Si, 7.1 g for Ba0.3Co4Sb12) was milled in a WC mill jar with two 11.2 mm WC spheres and four 7.9 mm WC spheres, sealed with Viton gaskets (WC mill jar set 8004, WC media 5004A, Viton gasket 39322, SPEX Sample Prep, Metuchen, NJ). The mill jar 24 was filled and sealed in an argon glove box, then the cap was wrapped with electrical tape (Super 88, 3m, St. Paul, MN), and a layer of Parafilm “M” wrapped around the entire jar (PM-996 Pechiney Plastic Packaging, Menasha WI) to prevent air from entering the jar. The powder, sealed in the mill jar with the media, was removed from the glove box, then milled for 10 minutes in a vibratory mill (8000M SPEX SamplePrep). The sealed mill jar was returned to the glove box after milling and opened to collect the powder. The oxygen concentration in the glove box did not change when the mill jar was opened, with a meter sensitivity of ± 0.1 ppm, indicating the mill jar remained sealed during milling. The milled powder was collected and stored in glass scintillation vials. For the YbAl3 sieved powder, the powder was VM in a stainless steel jar with stainless steel media. The powder was loaded into the jar in air, thus no electrical tape or parafilm was necessary to seal the jar. Cleaning of the VM jar and media was performed when changing the TE material milled, or when the jar was opened while removed from the glove box. For cleaning, the jar and media were filled to approximately 1 cm depth with play sand and milled for 10 minutes. The milled sand was removed, the jar disassembled and rinsed with RO water and ethanol, and the Viton gaskets wiped down with a Kim-wipe (Kimberly Clark 34155, Neenah, Wisconsin) and ethanol. The jar and gaskets were reassembled, the jar filled with approximately 1 cm of play sand, then with ethanol to the top of the sand and milled for a second 10 minutes. The jar was cleaned out, rinsed and gaskets wiped down again before reassembly. The jar was filled with approximately 1 cm of ethanol and milled for a third 10 minutes. The jar was cleaned out, rinsed and gaskets wiped down a third time before final reassembly. A separate set of Viton gaskets were used for each material 25 system milled in the jar. The Viton gaskets were changed during final reassembly if necessary. 2.4 Incorporation of nanoparticles Silicon carbide nanoparticles (SiCNP) from one of two vendors, with vendorspecified average particle size between 45 nm and 55 nm (44646, Lot number E15T018, Alfa Aesar) or 50 nm to 60 nm (4621HW, Lot #4621-110209, Nanostructured & Amorphous Materials Inc., Houston, TX) were incorporated into PM powder of SnTe (chapter 7) and Mg2Si (chapter 4) by PM at 120 RPM for 3 h, and into CGSR powder of Mg2Si by VM for 10 min. Other than the addition of SiCNP, and the change in PM speed and time, all PM and VM processing was identical to the processing Silver nanoparticles (AgNP) with a vendor specified purity of 99.9% and a vendorspecified size range of from 20 nm to 40 nm (45509, Alfa Aesar, Ward Hill MA), were dispersed into the Ba0.3Co4Sb12 powder by PM at 300 rpm for 15 min (chapter 6). 2.5 Sintering of powders Sintering of powders to make bulk specimens was performed by either hot press (HP) or by pulsed electric current sintering (PECS). 2.5.1 Hot press sintering Hot press (HP) sintering was employed for the sieved YbAl3 (chapter 3) and PM Ba0.3Co4Sb12 powder (chapter 6). HP Sintering of the Ba0.3Co4Sb12 powder (chapter 6) was performed in an argon glove box with a graphite die and a custom-built hot press. The graphite die was lined with grafoil (Graftech International, Lakewood, OH) to prevent the powders from sticking to the die. The temperature was monitored by K-type thermocouple and adjusted 26 manually by varying power supplied to a resistance coil heating element in a spiral surrounding the die. A manual hydraulic press was maintained at the desired pressure during the HP cycle. HP specimens of YbAl3 (chapter 3) were pressed by Edward J. Timm in a 22 mm diameter graphite die at 973 K for 1 hour under at 70 MPa (Hot press model HP20014020-23G by Thermal Technologies LLC, Santa Rosa, CA). 2.5.2 Pulsed electric current sintering Pulsed electric current sintering (PECS) was employed for the VM YbAl3 powder (chapter 3), CGSR, PM and VM Mg2Si powders (chapters 4 and 5), VM Ba0.3Co4Sb12 powder (chapter 6), PM SnTe powder (chapter 7), and the CGSR PbS and PbSe powders (chapter 8). Specimens were densified by pulsed electric current sintering (SPS Model 10-3, Thermal Technology LLC, Santa Rosa, CA) in a 12.7 mm or 20 mm diameter graphite die lined with grafoil. Powder was added to the die in an argon glove box and manually tamped down with the punch of the die. The loaded die was transferred to the PECS machine, and then the PECS chamber was evacuated with a roughing pump. The PECS chamber was backfilled with argon, then the sintering cycle was run from automated control software for the desired temperature and pressure ramp rate and hold time as specified for the individual specimens (see chapters 3-8). After cooling, the die was removed from the PECS chamber and the specimen removed. Overhanging grafoil may be cut with a razor blade to aid in removal of the specimen from the die. 27 2.6 Sintered specimen preparation After sintering, the specimens remained covered with a grafoil layer requiring removal. The bulk of the grafoil was removed with a #10 or #11 scalpel (372610 or 372611, Aspen Surgical Products, Caledonia, MI) by hand. The remaining grafoil on a rough sintered specimen surface was primarily removed by hand sanding until most of the rough surface was no longer visible using 600 to 1200 grit sand paper with adhesive back, adhered to an aluminum plate. For the Mg2Si specimens, the sand paper was wetted with ethanol due to an incompatibility between the Mg2Si and water. For all other specimens, the sand paper was wetted with RO water. Final cleaning of grafoil was performed with a Pyrex 50 mL beaker filled with 20 to 30 mL of ethanol (Mg2Si specimens) or a solution of 1% liquid detergent (Liquinox detergent, Alconox, White Plains, NY) in RO water. The beaker was placed in an ultrasonic cleaner (Ultramet III, Buehler, Evanston, IL) until the grafoil was removed, 5 to 30 minutes. 2.7 Mounting of specimens Specimens were mounted for cutting or polishing by thermoplastic or by epoxy. 2.7.1 Thermoplastic mounting of specimens Specimens were mounted with thermoplastic (CrystalBond 509, EMS, Hatfield, PA) to a glass slide for cutting or to aluminum stubs for polishing. The glass slide or stub, pieces of thermoplastic and specimen were heated simultaneously, starting with a room-temperature hot plate, to the softening point of the thermoplastic over 5 to 30 minutes. The slow heating rate was used to avoid introducing a thermal shock to the specimen. The specimen was placed into the softened thermoplastic such that the desired 28 side was flat against the slide or stub with only a thin layer of thermoplastic between, and the thermoplastic extended at least half the distance up the sides of the specimen. Additional small pieces of thermoplastic may be added if necessary to reach the required distance for specimen support. The heat was removed and the mounted specimen air cooled to room temperature. 2.7.2 Thermoplastic dismounting of specimens Dismounting of thermoplastic mounted specimens was performed by slowly reheating the mounted specimen on a hot plate, then removing the softened thermoplastic from the specimen with a metal spatula. The roughly cleaned specimen was slowly cooled to room temperature. Final cleaning of the specimen was done with acetone in a fume hood. A KimWipe with acetone was used to remove the bulk of the thermoplastic film by hand. The specimen was then soaked in approximately 10 to 20 mL of acetone in a 50 mL Pyrex beaker for approximately 5 minutes. The specimen was removed from the soak, rinsed with a spray of acetone while held in a clean KimWipe, then soaked in a second 50 mL Pyrex beaker of approximately 10 to 20 mL clean acetone for another 5 minutes. The specimen was removed from the second soak, rinsed with a spray of acetone while held in a clean KimWipe, then air dried on a clean and dry KimWipe. If any film of thermoplastic was observed, the final cleaning was repeated. 2.7.3 Epoxy mounting of specimens Small or irregular pieces of specimens for polishing and indentation were mounted in epoxy. Ring forms (Black Bakelite Ring Forms #811-221, Leco, St Joseph, MI) were placed on an aluminum plate that had been sprayed with mold release (3470 29 Reliable Release, Crown, Woodstock, IL). Specimens for mounting were placed in each ring form. The manufacturer specified ratio of epoxy resin and hardener (Epoxicure Resin and Epoxicure Hardener, Buehler, Evanston, IL) were measured by weight on an electronic balance (Adventurer AR2140, Ohaus Corp, Pine Brook, NJ) in a disposable medicine cup to produce 20-30 mL of epoxy. The resin and hardener were mixed by hand with a wooden stir stick until swirls of separate resin or hardener were no longer visible. The ring forms were filled to approximately 2/3 full with the mixed epoxy, and any bubbles that formed around the specimen were dislodged with the stir stick to dissipate at the surface. A steel weight was sprayed with mold release and placed on top of the ring forms to limit epoxy seepage out from the bottom of the ring forms. The epoxy was cured for at least 1 day. After curing, the epoxy was engraved with the specimen name using a rotary tool. Any epoxy that leaked out was cut from the sides of the ring forms. 2.8 Specimen cutting Specimens mounted on a glass slide were cut on a low speed diamond saw (Isomet Low Speed Saw, Buehler, Evanston, IL) with a cutting oil bath (VP-50 cutting oil, part number 811-024, Leco Corporation, St. Joseph, MI) set at a cutting speed between 3 and 4. Counterweights were adjusted to cut through a 10 mm specimen in approximately 15 to 30 minutes. 2.9 Specimen polishing Mounted specimens were polished on an automatic polishing machine (Leco Vari/Pol VP-50, Leco Corporation, St. Joseph, MI). Specimens were mounted in a 12position specimen holder, with 3 or 4 specimens polished at a time. A series of diamond 30 grit pastes (Table 2.1) were used to polish the specimens prior to indentation or scanning electron microscope (SEM) imaging of polished surfaces. The diamond grit was applied in a spiral dot pattern to a WhiteTec polishing pad (White Tec #812-454, Leco, St. Joseph, MI) mounted to a 12-inch polishing wheel. The wheel was wetted with diamond compound extender (Microid Diamond Compound Extender #811-004, Leco, St. Joseph, MI). The specimens were polished for 3 to 120 minutes at each grit size, until the specimens were flat and no scratches from the previous grit size were present (e.g., polishing with 6 µm grit is complete when no scratches are visible that were present after polishing with 9 µm grit). When polishing is complete for a specific grit size, the specimens and the mounting wheel were cleaned to remove all polishing grit prior to the next polishing step. All specimens except Mg2Si were cleaned by rinsing with tap water, then cleaning in an ultrasonic cleaner (B3500A-MT, VWR International, West Chester, PA) with a solution of 1% detergent (Liquinox detergent, Alconox) in RO water for 10 minutes. The specimens were then manually cleaned with cotton tipped applicators and ethanol before final rinse with RO water. For the Mg2Si specimens, the cleaning was performed exclusively by hand with cotton tipped applicators and ethanol with an ethanol rinse due to a reaction between Mg2Si and water. 31 Table 2.1. A series of diamond grit pastes were used to polish the specimens. A grit size may be skipped, with additional polishing time on the next smaller diamond grit size to compensate. Nominal Diamond Grit Manufacturer Part Number Diamond Grit Size Range Size 67 µm Warren Superabrasives, 54-80MB MUS 20gm 54-80 m Anaheim, CA 35 µm Not listed Warren Diamond Powder #35 MUS MB 20G Company, Olyphant, PA 9 µm Not listed Leco Corporation, 810-913 St. Joseph, MI 6 µm Not listed Warren Diamond Powder #6 MUS MB 20G Company, Olyphant, PA 1 µm Not listed Leco Corporation, 810-870 St. Joseph, MI 0.5 µm Not listed Leco Corporation, 810-868 St. Joseph, MI 32 2.10 Mass, dimensions and density Specimens of cylinder or parallelepiped geometry were measured to determine mass and dimensions, and calculate density. The mass and dimensions were used in later resonant ultrasound spectroscopy measurements (see section 2.11) and density was used in the analysis of mechanical property measurements. Each of the dimensions of a specimen were determined by 5 measurements with micrometers (293-832, Mitutoyo, Japan), one measurement on each corner and one in the center or similar pattern, and the 5 measurements averaged. Specimen mass was measured by electronic balance (Adventurer AR2140, OHAUS, Pine Brook IL). 2.11 Resonant ultrasound spectroscopy The Young’s modulus, E, shear modulus, G, and Poisson’s ratio, ν, of the specimens were determined by resonant ultrasound spectroscopy (RUS). In RUS, the resonant frequencies of a specimen are measured across a range of frequencies. These measured resonant frequencies are a function of the specimen mass, dimensions, geometry and elastic moduli [Migliori 1997; Ren 2009a]. The resonant frequencies were measured with a commercial RUS system (RUSpec, Quasar International, Albuquerque, NM) by placing the specimen on either a tripod arrangement of transducers (Figure 2.1) or between two transducers across a body diagonal. The range of frequencies was set to begin measurement before the first resonance frequency, and record at least 40 resonance frequencies (Figure 2.2). Four measurements of the resonance frequency spectrum were taken per specimen, two per side for the tripod arrangement, and once for each body diagonal for the two transducer arrangement. Some resonance frequencies may be missing or poorly measured for one 33 position, possibly due to contact issues between the specimen and the transducer or other causes [Ren 2009b], but more visible in another scan. The scan with the clearest representation of resonance frequencies was chosen for analysis. Analysis was performed with commercial software (Figure 2.3) for either a parallelepiped geometry specimen (RPModel version 2.68b, Quasar International) or a disk shaped specimen (CylModel version 2.68b, Quasar International). The measured mass and dimensions, and an initial guess for the stiffness, c11 and c44 for isotropic specimens, were entered, the desired number of calculated resonance frequencies, the convergence rate (typically set to 0.5), the polynomial order (typically set to 12), and the fit dimensions option set (typically unused). The calculated resonance frequencies were matched to measured resonance frequencies from the spectrum (Figure 2.2), and the model was fit to the results through an iterative process. Typically, at least 12 peaks were included on the first fit of the model. Additional peaks may be added or removed from the analysis and rerun the model fit to improve the output results. For elastically isotropic specimens, the elastic moduli are calculated from the stiffness as [Ren 2009b] ( ) (2.1) (2.2) ( (2.3) ) The output of the RUS model includes information on the confidence of fit. From the output information, the uncertainties dc11 and dc44 in the measurement of c11 and c44 correspond to the 95% confidence levels calculated by the RUS model software [Ren 34 2009b]. The dc11 and dc44 values are computed from the first and last values, A and D, under the “chisquare increased 2% by…” section of the RUS output file, as indicated in Figure 2.4. The dc11 and dc44 values are used to compute dE, dG, and dν. (2.4) (2.5) √( ( ) ) ( ) (2.6) (2.7) √ ( (2.8) ) The output of equations (2.6-2.8) are reported as the error for the RUS measurements of E, G and ν, respectively. Additional details of the RUS theory are provided in a comprehensive book by Migliori and Sarrao [Migliori 1997]. 35 Intensity (A.U.) Figure 2.1. Typical setup of tripod arrangement of RUS transducers, with specimen on top and the transducer tips near the edge of the specimen. 0 100 200 300 400 500 600 Frequency (kHz) Figure 2.2. RUS spectrum from PbSe specimen. Each sharp peak in intensity represents a mechanical resonance at the driven frequency. 36 Figure 2.3. RUS analysis software, showing setup of initial conditions for a PbSe specimen. 37 Cylinder Calculation Report This Report was prepared on 12/18/12 14:35:52 Eastern Standard Time PbSe Elastic Moduli Analysis is completed. free moduli are c11 and c44 Poly Order = 12, Mass = 0.639300 gm, Density = 7.647000 g/cc Convergence Rate = 0.500000 n fmeas(kHz) fcalc(kHz) %err 1 65.404 65.227 -0.27% 2 65.404 65.227 -0.27% 3 100.989 101.377 0.38% 4 131.858 132.144 0.22% 5 131.858 132.144 0.22% 6 159.460 160.289 0.52% 7 159.460 160.289 0.52% 8 183.645 184.620 0.53% 9 183.645 184.620 0.53% 10 186.862 191.681 2.58% ... Bulk Modulus= 38.252000 GPa c11 67.76 c22 67.76 c33 67.76 c23 23.50 wt 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 0.0 k 6 4 6 2 8 3 5 7 1 2 i 2 1 3 1 1 1 1 2 2 2 c13 23.50 c12 23.50 df/d(moduli) 0.05 0.95 0.05 0.95 0.59 0.41 0.07 0.93 0.07 0.93 0.01 0.99 0.01 0.99 0.19 0.81 0.19 0.81 0.41 0.59 c44 22.13 c55 22.13 c66 22.13 Young's Modulus= 55.657217 GPa Shear Modulus= 22.130000 GPa Poisson's Ratio = 0.257506 Longitudinal Velocity (c11)= 2.976742 mm/us Shear Velocity (c44)= 1.701160 mm/us Xdiam= 0.792070 cm, Ydiam= 0.792070 cm, Height= 0.169670 cm loop# 6, rms error= 0.3210%, changed by length of gradient vector= eigenvalues 1.15562 211.91354 0.000006 blamb= -0.0000005% 0.000000 eigenvectors 1.00 0.05 -0.05 1.00 chisquare increased 2% by the following % changes in independent parameters 0.47 -0.07 0.00 0.11 Figure 2.4. Sample output file from a CylModel RUS analysis, truncated to show only the first 10 resonance peaks. Note that resonance frequency 10, 191.681 kHz, was not included in the analysis to improve the error. The “chisquare increased 2% by…” section includes the highlighted values of A, 0.47, and of D, 0.11, used in the determination of the uncertainties, dc11 and dc44. 38 2.12 Hardness and fracture toughness by Vickers indentation Vickers indentations were performed with a Buehler microindenter (Buehler Semimacro Indenter, Lake Bluff, IL) and a Shimadzu hardness tester (Shimadzu HMV2000, Kyoto, Japan). The indenter machines were calibrated by use of a steel standard (761-048, Yamamoto Scientific Tools Lab, Co LTD, Japan), and a correction factor, ζ, of 0.95 to 1.00 for the Shimadzu hardness tester and 0.99 to 1.02 for the Buehler microindeter was calculated. The correction factor was multiplied by the measured hardness values for the specimens. Vickers indentations typically produce an indentation impression as well as a set of radial cracks. If cracks are present, the dimensions of the cracks are the largest feature size of the indentation, and if cracks are not present, the dimensions of the indentation impression are the largest feature. Indentations were separated from the edges of the specimen and from other indentations by a minimum of 5x the largest feature of the indentation. This separation distance is inclusive of a minimum thickness for the specimen. Typically, the indentations were 1.0 mm from any edge and 0.5 mm from the nearest indentation. Hardness was calculated from the Vickers indentation impression (Figure 2.5), 2a, ( ) (2.9) where P is the applied load, ζ is the indenter correction factor, and 1.8544 is a geometric factor for the angle of the pyramidal Vickers indenter tip [Wachtman 2009]. 39 Radial cracks from Vickers indentations, 2c (Figure 2.5), may be used to estimate the fracture toughness, KIC. As Vickers indentations is not purely mode one fracture toughness measurement, but rather an estimate, the designation of KC shall be used when referencing fracture toughness in general rather than mode one fracture toughness, and for values of fracture toughness determined by Vickers indentation. The KC was determined by [Wachtman 2009] ( ⁄ ) ⁄ (2.10) ⁄ where ξ is a calibration factor set to 0.0016 [Anstis 1981], E is the Young’s modulus for the specimen as determined by RUS analysis (see section 2.11), and P is the applied load. The indentation method is not without controversy, with the assumption of the crack shape and plastically deformed zone being both revised and expanded [Anstis 1981; Dukino 2006] and questioned [Eilertsen 2013; Quinn 2007], with Quinn and Bradt stating, “the VIF technique is not suitable for the measurement of the fracture toughness, KIc, or any other form of the fracture resistance of ceramics or other brittle materials” [Quinn 2007]. It should be noted that some of the indentations used to question the indentation method in Eilertson et al. [Eilertsen 2013] have crack systems that are not suitable for determination of KC by Vickers indentation, as illustrated by Anstis et al. [Anstis 1981], and simply illustrate that the use of poorly developed radial cracks produce poor results for KC. Regardless, the Vickers indentation method for determination of KC is a common test with particular value for specimens where other methods of measuring KC are not possible or practical [Wachtman 2009; Walker 2011; Ni 2010; Zhao 2008; Anstis 1981; Evans 1976]. 40 2c 2a Figure 2.5. Schematic representation of Vickers indentation, illustrating the indentation impression dimensions, 2a, and the radial crack length dimensions, 2c. 41 2.13 X-ray diffraction X-ray diffraction (XRD) analysis was performed on YbAl3 (chapter 3), Mg2Si (chapter 4), SnTe (chapter 7), and PbS (chapter 8) to verify phase, determine lattice parameter, and detect possible contamination. Diffraction was performed with a benchtop XRD machine (Miniflex II, Cu Kα radiation, Rigaku, Tokyo, Japan) on powder SnTe and powder YbAl3, or with a rotating stage free-standing XRD machine (Bruker Davinci Diffractometer) at the Michigan State University Center for Crystallographic Research on powder Mg2Si and sintered PbS. Both XRD machines use a Cu Kα radiation source. Reitveld refinement and phase determination was performed with commercial software (HighScorePlus version 3.0c, PANalytical B.V., Almelo, The Netherlands) and patterns from the crystallography open database [Gražulis 2012; Gražulis 2009]. 2.14 Scanning electron microscopy imaging and energy dispersive x-ray spectroscopy Scanning electron microscopy (SEM) imaging permits the examination of micrometer- and nanometer-scale features such as determining size or morphology of particles, pores, and grains, determination of differences in average atomic weight by backscatter imaging, and determining distribution of particles, pores, or grains. For these reasons, SEM imaging was performed on each of the six studies in this dissertation. The SEMs used in this study (6400, 6610LV, and 7500F, JEOL Ltd., Japan) at the Center for Advanced Microscopy, Michigan State University were operated at an average working distance of 15 mm (6400 and 6610LV) or 8 mm (7500F) and an accelerating voltage of 15 kV, unless specified. 42 Powder specimens were mounted with carbon paint (05006-AB, SPI Supplies, West Chester, PA), and sintered specimens were mounted with epoxy and carbon paint, or with carbon tape. Prior to imaging, each specimen was blown off with compressed air (22531093, OfficeMax) to remove loose powder or dust on the surface. Energy dispersive x-ray spectroscopy (EDS) was performed primarily with the 6610LV SEM because the tungsten filament and detector are capable of recording higher counts than the 7500F SEM. The 7500F SEM was used for EDS on only the PM-HP673-Ag specimen (chapter 6) because the higher resolution of the 7500F SEM permitted locating the necessary areas for EDS detection more readily than was possible on the 6610LV. 2.14.1 Grain size measurement from fractured surface images Fractured surfaces were used to determine average grain sizes by the linear intercept method [ASTM-Standard-E112-13 2014] in specimens with unimodal grain size distributions. Specimens with bimodal grain size distributions are not appropriate for measurement by lineal intercept method. A fractured surface image with approximately 10-30 grains along the length and clearly visible grain boundaries was chosen for analysis. In the linear intercept method, a line of measured length is drawn in a random orientation across a fractured surface image, and the number of grain boundaries the line intercepts was summed. Additional lines were drawn until the number of intercepts counted was greater than 200, typically about 10 to 15 lines. The total length of the lines was scaled according to the printed scale bar on the image, and divided by the total number of intercepts to determine an average line length between intercepts. The average 43 line length was multiplied by the stereographic projection factor, 1.5, to determine the average grain size of the specimen. 44 REFERENCES 45 REFERENCES [Anstis, Chantikul, Lawn, Marshall 1981] Anstis, G., Chantikul, P., Lawn, B., Marshall, D. A critical evaluation of indentation techniques for measuring fracture toughness: I, direct crack measurements. Journal of the American Ceramic Society 64 (1981) 533–538. [ASTM-Standard-E112-13 2014] ASTM-Standard-E112-13. Standard Test Methods for Determining Average Grain Size. In: ASTM Volume 03.01 Metals Mechanical Testing; Elevated and Low Temperature Tests; Metallography. ASTM International, West Conshohocken, PA, 2014. [Dukino, Swain 2006] Dukino, R.D., Swain, M. V. Comparative Measurement of Indentation Fracture Toughness with Berkovich and Vickers Indenters. 2 (2006) 3299–3304. [Eilertsen, Subramanian, Kruzic 2013] Eilertsen, J., Subramanian, M. a., Kruzic, J.J. Fracture toughness of Co4Sb12 and In0.1Co4Sb12 thermoelectric skutterudites evaluated by three methods. Journal of Alloys and Compounds 552 (2013) 492–498. [Evans, Charles 1976] Evans, A., Charles, E. Fracture toughness determinations by indentation. Journal of the American Ceramic …(1976) 371–372. [Gražulis, Chateigner, Downs, Yokochi, Quirós, Lutterotti, Manakova, Butkus, Moeck, Le Bail 2009] Gražulis, S., Chateigner, D., Downs, R.T., Yokochi, a F.T., Quirós, M., Lutterotti, L., Manakova, E., Butkus, J., Moeck, P., Le Bail, A. Crystallography Open Database - an open-access collection of crystal structures. Journal of applied crystallography 42 (2009) 726–729. [Gražulis, Daškevič, Merkys, Chateigner, Lutterotti, Quirós, Serebryanaya, Moeck, Downs, Le Bail 2012] Gražulis, S., Daškevič, A., Merkys, A., Chateigner, D., Lutterotti, L., Quirós, M., Serebryanaya, N.R., Moeck, P., Downs, R.T., Le Bail, A. Crystallography Open Database (COD): an open-access collection of crystal structures and platform for world-wide collaboration. Nucleic acids research 40 (2012) D420–7. [Migliori, Sarrao 1997] Migliori, A., Sarrao, J.L. Resonant Ultrasound Spectroscopy: Applications to Physics, Materials Measurements, and Nondestructive Evaluation. Wiley-Vch, New York, 1997. [Ni, Case, Khabir, Stewart, Wu, Hogan, Timm, Girard, Kanatzidis 2010] Ni, J.E., Case, E.D., Khabir, K.N., Stewart, R.C., Wu, C.-I., Hogan, T.P., Timm, E.J., Girard, S.N., Kanatzidis, M.G. Room temperature Young’s modulus, shear modulus, Poisson's 46 ratio and hardness of PbTe–PbS thermoelectric materials. Materials Science and Engineering: B 170 (2010) 58–66. [Pilchak, Ren, Case, Timm, Schock, Wu, Hogan 2007] Pilchak, A.L., Ren, F., Case, E.D., Timm, E.J., Schock, H.J., Wu, C.-I., Hogan, T.P. Characterization of dry milled powders of LAST (lead–antimony–silver–tellurium) thermoelectric material. Philosophical Magazine 87 (2007) 4567–4591. [Quinn, Bradt 2007] Quinn, G.D., Bradt, R.C. On the Vickers Indentation Fracture Toughness Test. Journal of the American Ceramic Society 90 (2007) 673–680. [Ren, Case, Ni, Timm, Lara-Curzio, Trejo, Lin, Kanatzidis 2009a] Ren, F., Case, E.D., Ni, J.E., Timm, E.J., Lara-Curzio, E., Trejo, R.M., Lin, C.-H., Kanatzidis, M.G. Temperature-dependent elastic moduli of lead telluride-based thermoelectric materials. Philosophical Magazine 89 (2009) 143–167. [Ren, Case, Ni, Timm, Lara-Curzio, Trejo, Lin, Kanatzidis 2009b] Ren, F., Case, E.D., Ni, J.E., Timm, E.J., Lara-Curzio, E., Trejo, R.M., Lin, C.-H., Kanatzidis, M.G. Temperature-dependent elastic moduli of lead telluride-based thermoelectric materials. Philosophical Magazine 89 (2009) 143–167. [Schmidt, Case, Ni, Sakamoto, Trejo, Lara-curzio, Payzant, Kirkham 2011] Schmidt, R., Case, E., Ni, J., Sakamoto, J., Trejo, R., Lara-curzio, E., Payzant, E., Kirkham, M. Characterization of the mechanical properties of two n-type and one p-type Sb-based skutterudite thermoelectric compounds. 2011. [Wachtman, Cannon, Matthewson 2009] Wachtman, J.B., Cannon, W.R., Matthewson, M.J. Mechanical Properties of Ceramics. Wiley-Vch, Hoboken, N.J., 2009. [Walker, Marotto, Rafiee, Koratkar, Corral 2011] Walker, L.S., Marotto, V.R., Rafiee, M. a, Koratkar, N., Corral, E.L. Toughening in graphene ceramic composites. ACS nano 5 (2011) 3182–90. [Zhao, Zhang, Li, Zhou, Liu, Liu 2008] Zhao, L.-D., Zhang, B.-P., Li, J.-F., Zhou, M., Liu, W.-S., Liu, J. Thermoelectric and mechanical properties of nano-SiC-dispersed Bi2Te3 fabricated by mechanical alloying and spark plasma sintering. Journal of Alloys and Compounds 455 (2008) 259–264. 47 3 Room temperature mechanical properties of polycrystalline YbAl3, a promising low temperature thermoelectric material Robert D. Schmidt, Eldon D. Case, Gloria J. Lehr, Donald T. Morelli Chemical Engineering and Materials Science Department, Michigan State University, East Lansing, MI, 48824 Published in: Intermetallics 35 (2013) 15-24. Abstract Intermetallic YbAl3 in the L12 (AuCu3) phase is a promising material for low temperature thermoelectric applications. However, there is no experimental data in the literature on the mechanical properties of YbAl3, although the design and development of thermoelectric modules incorporating YbAl3will require mechanical property data. Using resonant ultrasound spectroscopy (RUS), the room temperature Young’s modulus, shear modulus, bulk modulus and Poisson’s ratio were determined as a function of volume fraction porosity, P, for specimens densified by both hot pressing (HP) and pulsed electric current sintering (PECS) polycrystalline specimens, where P ranged from 0.030 to 0.233 and mean grain sizes ranged from 0.5 to 1.4 µm. In addition, the longitudinal and acoustic wave speeds and the Debye temperature were measured. Using Vickers indentation, the hardness and fracture toughness of the specimens were also measured. Despite microstructural differences between the HP and PECS-processed specimens, the porosity dependence of the mechanical properties was a function of the total volume fraction porosity, P, independent of the details of the size and spatial distribution of pores within individual specimens. Keywords: A. aluminides, B. elastic properties, mechanical properties at ambient temperature 48 3.1 Introduction Research directed at developing more efficient thermoelectric (TE) materials has surged in the last decade with hopes that applications in thermoelectric power generation may be realized on a large scale. TE devices convert heat directly to electricity using the Seebeck effect. The efficiency of conversion depends on the dimensionless thermoelectric figure of merit: ZT   2 PF T T   (1) In this equation, α is the Seebeck coefficient, σ the electrical conductivity, and κ the thermal conductivity of the material; PF is termed the power factor. Historically, the best TE materials have exhibited ZT values not exceeding unity. Recent research has successfully demonstrated higher values of the figure of merit, most notably in bulk nanostructured chalcogenides [Biswas 2011] and filled skutterudite [Shi 2011] compounds, which exhibit ZT values in the range of 1.6-1.8 above 600 K. Another application of thermoelectricity involves solid state heating and cooling using the Peltier effect. In contrast to the very promising developments in the high temperature range, new materials for Peltier cooling at room temperature and below have received relatively little attention. One interesting and promising material for Peltier cooling is the intermetallic compound YbAl3. This compound exhibits a power factor some five to ten times that of bismuth telluride [Van Daal 1974; Rowe 2002; Mahan 1998] (the state of the art material for TE cooling), but does not possess high ZT due to its high thermal conductivity. Thermal conductivity reduction in YbAl3-based materials is a current area of focus of our research. 49 While it is self-evident that the electronic and thermal transport properties of a material are important parameters in determining its suitability in a TE application, equally important, yet often underappreciated, is an understanding of the mechanical properties. By the very nature of their application, TE materials can be subjected to large thermal gradients, temperature cycling, and thermal shock. In this study, the room temperature mechanical properties (elasticity, hardness and fracture toughness) were measured as a function of volume fraction porosity, P, for polycrystalline intermetallic YbAl3 specimens in the L12 phase. The specimens were densified by either hot pressing (HP) or pulsed electric current sintering (PECS). Resonant Ultrasound Spectroscopy (RUS) was used to measure the room temperature elastic moduli, including Young’s modulus, shear modulus, bulk modulus and Poisson’s ratio. The elastic moduli are required for stress and strain modeling of any solid which in turn is critical in the design of thermoelectric modules that will be subjected to both thermal gradients and thermal transients. Also, the longitudinal and acoustic wave speeds of YbAl3 were measured by RUS. The wave speeds were in turn used to compute the acoustic Debye temperature which is related to thermal conductivity. Hardness, which is a measure of the resistance of a material to plastic deformation and scratching, was measured by Vickers indentation. Finally, the fracture toughness, KC, was estimated using the length of radial cracks induced by Vickers indentation, where KC measures the resistance to slow crack growth. The mechanical properties were characterized as a function of porosity for two reasons. First, extrapolation of the experimental modulus versus temperature data to P = 0 allows the direct comparison to theoretical modulus calculations by techniques such as 50 density functional theory. Second, determining the modulus-porosity relationship allows comparison of the data from this study to other experimental studies by other researchers for specimens with differing P values. 3.2 Experimental procedure 3.2.1 Materials and specimen preparation Only two binary phases exist in the Yb-Al phase diagram [Massalski 1987]: YbAl2 and YbAl3. YbAl3 can be difficult to synthesize in single phase form due to the peritectic nature of the phase transformation at the 1:3 Yb:Al ratio. Samples which are arc-melted at this composition invariably contain mixtures of YbAl3, Al, and the highmelting point YbAl2 phase which cannot be removed even after extended annealing. Single crystals of pure YbAl3 have been synthesized by growth in an Al flux [Canfield 1992]; the excess Al prevents the formation of the YbAl2 phase. Unfortunately, the crystals so obtained are too small for mechanical property measurements. We have successfully synthesized pellets of YbAl3 of up to 2.54 cm in diameter by a combination of a novel arc-melting procedure and either HP or PECS. In order to obtain pure YbAl3 phase material, we applied the concept of growth of YbAl3 in an Al flux to the arc-melting process. Starting materials (Al 99.999% - ULVAC and Yb 99.9% - Alfa Aesar) were arc melted on a water-cooled copper hearth under argon with the nominal composition YbAl7. Each ingot was melted and flipped at least 5 times to ensure homogeneity. This process resulted in a polycrystalline ingot comprised of a mixture of YbAl3 and Al phases, as determined by X-ray diffraction (Rigaku MiniFlex II with Cu Kα radiation). Just as in the growth of single crystals in an Al flux, the excess Al present in 51 the arc-melting process inhibits the formation of the undesirable YbAl2 phase. To remove the pure Al phase, the ingots were broken into small pieces and etched in a 5M NaOH bath for 24 hours. The pieces were then rinsed in a series of H2O baths to remove any remnant NaOH, heated on a hotplate for 10-15 minutes to dry, and then ground into powder using a mortar and pestle. The powder was then sieved through a 100-mesh screen. Particles larger than this were returned to the NaOH bath and the process was repeated. The sieved powder was ball milled in a vibratory mill using stainless steel media. Prior to densification, the powder so obtained was determined by x-ray diffraction to be single phase YbAl3 (cubic AuCu3 structure with lattice constant of 0.4205 nm, in agreement with literature values [Rowe 2002; Havinga 1973]). The powder was then sieved again through a 100-mesh screen. Two densification techniques were used, HP and PECS, and the processing conditions are summarized in Table 3.1. The HP specimens were pressed in a 22 mm diameter graphite die at 973 K for 1 hour under at 70 MPa (Hot press model HP200-14020-23G by Thermal Technologies LLC, Santa Rosa, CA). The PECS samples were pressed in a 12.7 mm diameter graphite die at 1023 K to 1073 K for 1 hour at a pressure of 60 MPa (Model 10, Series 4 by Thermal Technologies LLC, Santa Rosa, CA). X-ray diffraction on the densified pellets again revealed predominantly single phase material. Several low intensity peaks, however, did not index to the AuCu3 phase, but rather matched fairly closely those of Yb3Al5O12, suggesting a slight amount of this phase exists as an impurity in the pressed pellets. 52 Table 3.1. The HP and PECS processing parameters for the YbAl3 specimens included in this study. Specimen YbAl3-A YbAl3-B YbAl3-C YbAl3-D YbAl3-E Densification technique HP HP PECS PECS PECS Sintering time (min) 75 60 60 10 10 Sintering temperature (K) 1053 1023 1073 1073 873 53 Pressure (MPa) 60 70 60 60 60 Specimen diameter (mm) 22 22 12.7 12.7 12.7 Although each of the YbAl3 specimens were fabricated as discs, prior to microstructural characterization and mechanical property measurement, selected specimens were cut to form rectangular bars and crescent- shaped specimens from the periphery of the bar. The crescent- shaped specimens were used in the microstructural examination and hardness testing while the RUS measurements were made on both the uncut disks and the as-cut bar-shaped specimens. Initially the specimens were cut using an electrical discharge machine, but due to difficulties in cutting the specimens, all subsequent cutting was done using a surface grinder. Prior to hardness and fracture toughness testing, the specimens were polished using diamond paste with a series of grit sizes ranging from 90 microns to 1 micron. 3.2.2 Microstructural characterization A scanning electron microscope (SEM, JEOL-6400, JEOL Ltd., Japan) with an accelerating voltage of 15 kV and a working distance of 15 mm was used to examine the size and morphology of the grains and pores for both the fractured surfaces and polished surfaces of the YbAl3 specimens. The electrical conductivity of the specimens was sufficiently high that no conductive surface coatings were required prior to SEM examination. A mean grain size was calculated using the linear intercept technique from SEM micrographs of the fractured surfaces of specimens. The volume fraction porosity was calculated from the specimen mass and dimensions. The mass of each specimen was measured by an electronic balance with a vendor-specified accuracy of ± 0.0003 g. The specimen dimensions were measured by a digital micrometer. 54 3.2.3 Hardness testing For the polished specimens, Vickers indentation was performed at loads from 2.94 N to 19.6 N (Shimadzu Model HMV-2000 indenter). The centers of each indentation impression were located at least 1 mm from the specimen edge, with the indentation impression centers spaced approximately 0.5 mm apart. The hardness, H, was calculated from [Wachtman 2009] H 1.8544 L 2a 2 (1) where L is the applied load and 2a is the average diagonal length of the indentation impression. The average hardness was calculated from the mean of at least 10 indentations on a given specimen. A steel calibration block with a hardness, H, of 7.75 GPa (761-048, Yamamoto Scientific Tools Laboratory, LTD, Japan) was used to calibrate the Vickers indenter. 3.2.4 Elastic modulus measurements The elastic moduli were measured using a commercial resonant ultrasound spectroscopy apparatus (RUSpec, Magnaflux Quasar, Albuquerque, NM), where the moduli were determined from the mechanical resonance frequencies, the mass, dimensions and geometry of the specimens [Migliori 1997, Ren 2008, Ren 2009]. Two specimen geometries were employed in this study, a disk and a rectangular bar. Since the governing equations for these two geometries differ, the moduli for the disk and bar-shaped specimens were calculated with different commercial software packages, namely CylModel (Magnaflux Quasar, Albuquerque, NM) for the disk 55 geometry and RPModel (Magnaflux Quasar, Albuquerque, NM) for the rectangular bar geometry. Two different transducer stages were used during the modulus measurements (Fig. 3.1a and b). For the tripod transducer stage (Fig. 3.1a), either a disk or bar-shaped specimen was set onto the three transducers, where one driver transducer induced the mechanical vibrations in the specimens and the remaining two transducers detected the specimen resonances (Fig. 3.1a). In the second type of transducer configuration, namely the two-transducer or bipod transducer stage, a bar-shaped specimen was placed along its face diagonal between a pair of opposing transducers, one driver transducer and one pickup transducer (Fig. 3.1b). For both the tripod and bipod transducer stages, the sinusoidal driving frequency was swept from 100 kHz to 600 kHz. The use of the two different specimen geometries and the two different transducer stages allowed independent measurements on a single specimen with a fixed modulus value, as will be discussed further in Section 3.3.3. 3.2.5 Fracture toughness measurements The fracture toughness, KC, was determined from the length of the Vickers indentation induced radial crack lengths using the relationship [Wachtman 2009] KC   ( E / H )1/ 2 L (3) c3/ 2 where ξ is a dimensionless calibration constant (0.0016) [Anstis 1981], L is the applied indentation load, and c is half of the radial crack length. In Eq. 3, H is the specimen’s hardness determined by Vickers indentation and E is the specimen’s Young’s modulus 56 determined by RUS. The mean and standard deviation of the fracture toughness were determined from at least 10 Vickers indentations per specimen. 3.3 Results and Discussion 3.3.1 Microstructural characterization The five YbAl3 specimens fabricated in this study had volume fraction porosity, P, ranging from 0.030 to 0.233 (Table 3.2). For each YbAl3 specimen in this study, the mean grain size ranged from about 0.5 to 1.4 µm (Table 3.3), thus there were no significant differences between the average grain sizes obtained by hot pressing and by PECS. However, SEM micrographs of fracture surfaces of each of the five specimens indicate that the size and spatial distribution of the HP and the PECS specimens differed from one another (Fig. 3.2). The microstructures of both hot pressed specimens YbAl3-A (Fig. 3.2a) and specimen YbAl3-B (Fig. 3.2b) consisted of (1) relatively dense “islands” with roughly micron grain sizes surrounded by (2) a more porous region with submicron average grain size. Two of the PECS specimens were relatively dense with small, uniformly distributed pores (Fig. 3.2c and d), while a PECS specimen intentionally processed to have higher porosity (P ¼ 0.23) than the other two PECS specimens again showed dense regions adjacent to more porous areas (Fig. 3.2e). The less dense regions in specimens YbAl3-A, YbAl3-B and YbAl3-E are marked by the ovals in Fig. 3.2a, b and e respectively. 57 Figure 3.1. Photographs of tripod and bipod transducer stages for the RUS measurements of elastic modulus performed in this study. Table 3.2. Specimen label, densification method (hot pressed, HP, or pulsed electric current sintering, PECS), mass, dimensions and mass density of each of the YbAl3 specimens included in this study. Specimen Densification Mass Dimensions (cm) Mass P label method (g) Density (g/cm3) YbAl3-A (bar) HP 1.0276 0.874 × 0.808 × 0.292 4.98 0.125 YbAl3-B (disc) HP 5.3646 Ø 2.213 × 0.320 4.36 0.233 YbAl3-B (bar) HP 0.9581 1.091 × 0.320 × 0.617 4.45 0.217 YbAl3-C (disc) PECS 1.8983 Ø 1.256 × 0.287 5.34 0.060 YbAl3-C (bar) PECS 0.8125 0.869 × 0.607 × 0.286 5.38 0.054 YbAl3-D (bar) PECS 0.8675 0.922 × 0.656 × 0.260 5.52 0.030 YbAl3-E (bar) PECS 1.2582 0.973 × 0.719 × 0.409 4.39 0.227 Table 3.3. The number of indentations, indentation load and mean hardness, , of the YbAl3 specimens and porosity, P. In each case, the loading time was 10 s. The porosity reported is the average as measured from the disc and bar specimen geometries. Specimen Number of Load Average volume GS label indentations (N) (GPa) fraction porosity 10 0.98 4.81 ± 0.73 0.125 YbAl3-A 20 2.94 4.86 ± 0.74 0.125 0.7 10 19.61 4.89 ± 0.39 0.125 YbAl3-B 10 0.98 2.80 ± 0.49 0.225 0.7 YbAl3-C 10 0.98 5.81 ± 0.15 0.057 1.4 YbAl3-D 10 0.98 6.81 ± 0.30 0.030 0.5 YbAl3-E 10 0.98 2.76 ± 0.66 0.227 1.2 58 Figure 3.2. SEM micrographs of fracture surfaces and polished surfaces of YbAl3 specimens included in this study. The ovals indicate more porous regions neighboring more dense areas that occur in the two HP specimens (Fig. 3.2a and b) as well as the PECS specimen intentionally processed to give high porosity (Fig. 3.2e). 59 Figure 3.3. The fractured surfaces of specimens YbAl3-B and YbAl3-E exhibited lower porosity regions surrounded by channels of higher porosity in (a) and (b), both with similar porosity but each sintered by a different method. In contrast, specimen YbAl3-D, shown in (c), exhibited a uniform density throughout. 60 The micrographs included in Fig. 3.3 were taken at a lower magnification than those in Fig. 3.2 in order to better illustrate differences in the spatial distribution of pores among the YbAl3 specimens included in this study. The fractured surfaces of hot pressed specimen YbAl3-B (P = 0.22) and PECS-processed specimen YbAl3-E (P = 0.23) exhibited relatively dense regions surrounded by porous regions, where the pores were either quasi-spherical or lenticular in shape with a wide size distribution from roughly sub-micron to 10 µm dimensions (Fig. 3.3a and b). Although the total volume fraction porosity, P, is very similar for specimens YbAl3-B and YbAl3-E, both the spatial distribution of pores and the densification technique were different for the two specimens. Note that for the hot pressed specimen YbAl3-B, the porous regions appear as bands roughly 100 µm or more across, while in the PECS processed specimen YbAl3E, the bands of porosity are roughly 20 µm across with a maximum pore size that is somewhat smaller than in YbAl3-B. Densifying by HP yielded specimens with volume fraction porosities of 0.125 (YbAl3-A) and 0.22 (YbAl3-B). In contrast, PECS processing yielded specimens with volume fraction porosities of 0.057 (YbAl3-C) and 0.030 (YbAl3-D). In order to produce an YbAl3 specimen with porosity comparable to the HP specimens, the PECS-processed specimen E was fabricated at a sintering temperature of 873 K for only 10 minutes rather than 1073 K for 60 minutes (Table 3.1). The differences in the processing techniques and the resulting microstructures for the HP and the PECS-processed specimens is quite significant in terms of the observed P dependence of the mechanical properties. As will be discussed in Sections 3.3.2 and 3.3.4, despite these processing/microstructural differences, the porosity dependencies of 61 the hardness and the elastic moduli are functions of the total volume porosity, P, only, rather than showing property differences based on the size or spatial distribution of porosity in individual specimens. 3.3.2 Hardness The decrease in H with increasing volume fraction porosity, P, is often expressed in terms of the empirical relationship H = H0 exp (– bHP) (4a) where bH is a measure of the rate of decrease in H with increasing porosity P and H0 is the value of hardness corresponding to a fully dense specimen [Rice 1998]. As reviewed by Rice [Rice 1998], for a number of oxides, non-oxides (including Si3N4, TiC, B4C, and glassy carbon) and the metals Cu and Fe, the decrease in H with increasing P can be described well by Eq. (4a) with bH values ranging from roughly 2 to 6. Furthermore, for specimens with low to intermediate values of P (such as the YbAl3 specimens included in this study), a Taylor series expansion of Eq. (4a) leads to the linear approximation H = H0 (1 – bHP) (4b) In this study, the linear form of the P versus H relationship (Eq. (4b)) describes the YbAl3 data well, as indicated by a least-squares fit of the hardness-porosity data to Eq. (4b) which yields a coefficient of determination, r2, of 0.965, a bH value of 2.65 + 0.18 and H0 = 6.99 + 0.15 GPa (Fig. 3.4). Thus despite the differences in processing and microstructure observed for the specimens in this study (Section 3.3.1, Figs. 3.2 and 3.3), the porosity dependence of H can be described well by a function of the total volume fracture porosity (Eq. (4b)), without regard to the details of the differing sizes or spatial distribution of pores. In addition, although no studies of the H versus P behavior of rare62 earth or transition metal trialuminides are available in the open literature, the bH value for YbAl3 found in this study falls within the range of the values typically reported in the literature for a broad range of solid materials. In addition to the lack of data on the H versus P behavior of trialuminides, no room temperature hardness data are available in the literature for a rare-earth or transition metal L12 trialuminide material. However, a room temperature Vickers indentation hardness value, H, of 5.0 + 0.5 GPa was measured by Milman et al. [Milman 2001] for cast ingots of TiAl3 (tetragonal D022 phase) using a 2.34 N load which is roughly comparable to the H value obtained in this study for PECS processed YbAl3 (cubic L12, Au3Cu form) of 6.81 ± 0.30 GPa at an indentation load of 0.98 N (Table 3.3). The load dependence of H for YbAl3 was also explored since in some cases H is a function of the applied indentation load [Nix 1998, Weaver 2003, Sangwal 2000, Milman 2011, Rice 1998], where with increasing indentation load the measured H can either increase (indentation size effect, ISE [Nix 1998, Weaver 2003, Milman 2011]) or decrease (reverse indentation size effect, RISE [Sangwal 2000]). The ISE can be especially pronounced in nanoindentation studies [Milman 2011]. In this study, the H value of YbAl3 is essentially independent of the indentation load over the range from 0.98 to 19.6 N (Table 3.3). Over load ranges similar to that used in this study, the measured hardness also has been reported to be relatively insensitive to the applied Vickers indentation load for other brittle thermoelectric materials including Mg2Si [Schmidt 2012], PbTe-PbS [Ni 2010] and LAST [Ren 2009]. 63 Hardness (GPa) 7 Hardness, PECS Hardness, HP 6 5 H0 4 3 2 0.0 0.1 0.2 0.3 P Figure 3.4. The hardness, H, versus volume fraction porosity, P, for the YbAl3 specimens included in this study. The solid line represents a least-squares fit to Eq. (4b). 64 3.3.3 Fracture Toughness For the specimens examined in this study, the fracture toughness, KC, by Vickers indentation of YbAl3 was between 1.13 MPa m1/2 and 1.69 MPa m1/2 (Table 3.4). No KC values for rare earth trialuminides are available in the literature. However there are the following fracture toughness data for the following single phase L12 transition metal trialuminides: (1) (Al +12.5 at.% Mn)3Zr, where KC increased from 1.86 MPa m1/2 to 2.16 MPa m1/2 during annealing at 800ºC to 1000ºC [Lee 2006], (2) (Al + x at.% Mn)3Ti, x = 3 to 12, with KC ranging from 1.1 MPa m1/2 to 1.54 MPa m1/2 as a function of x [Jang 2004], (3) (Al + 8 at.% Mn)3Ti, with a KC value of 1.54 MPa m1/2 which increased to 4.12 MPa m1/2 when annealed to 1100ºC [Jang 2004], and (4) Al3Ti(9Mn), with KC ≈ 2 MPa m0.5 for Vickers indentation loads from 200 g to 2000 g and a range of porosity from 0.005 to 0.12 [Varin 2001]. Thus, while the heat treatment and alloying of these transition metal trialuminides from the literature was primarily aimed at optimizing the mechanical properties, especially KC [Lee 2006; Jang 2004; Varin 2001], it is interesting that the KC values for the thermoelectric YbAl3 specimens in this study is roughly comparable to the KC measured for the transition metal trialuminides. 65 Table 3.4. For YbAl3 specimens included in this study, the fracture toughness, KC, measured by Vickers indentation, including the applied load, specimen grain size and volume fraction porosity, P. The KC was obtained from the mean of 10 indentations per specimen. Specimen YbAl3-A YbAl3-D KC (MPa m1/2) 1.69 ± 0.20 1.13 ± 0.07 Indentation load (N) 19.61 0.98 66 Grain size P (µm) 0.65 0.125 0.54 0.030 3.3.4 Elastic modulus As is the case for hardness, the P dependence of the elastic moduli is often expressed empirically in terms of an exponential decrease in moduli with increasing P which can be approximated by the linear functions E = E0(1 - bEP) (5a) G = G0(1 - bGP) (5b) B = B0(1 - bBP) (5c) where E is Young’s modulus, G is the shear modulus and B is the bulk modulus and E0, G0 and B0 are the P = 0 intercepts of the E, G, and B data, respectively (Table 3.5, Fig. 3.5a). The material dependent values of bE, bG and bB are measures of the decrease in the respective moduli as a function of increasing P. Rice [Rice 1998] also reviewed the P dependence of the elastic moduli E, G and B. The values of the coefficients bE, bG and bB (Eqs. (5a) – (5c)) typically range from about 2 to 6 for metals (Fe, Co, Ni, Cu, Be, Cu-Sn), polymer-based materials (epoxy and polyester) in addition to a number of oxide, carbide and nitride materials. For this study of the elasticity-P behavior of YbAl3, a least-squares fit of the modulus-porosity data to Eqs. (5a) – (5c), respectively (Fig. 3.5a), resulted in bE, bG and bB values that ranged from 2.21 + 0.17 (bB) to 2.38 + 0.05 (bG) (Table 3.5). Note that the bE, bG and bB coefficients obtained for YbAl3 in this study fall within the range of bE, bG and bB values reported by Rice [Rice 1998]. Thus, as was the case for the P dependence of H, the P dependence of the Young’s, shear and bulk moduli of YbAl3 is consistent with elasticity versus P relationships reported for a number of ceramic, polymer and metallic materials in the literature [Rice 1998]. Also in accord with the observed P dependence of H, the P 67 dependence of the elastic moduli are described well by functions of the total volume fraction porosity, P (Eqs. (5a) – (5c)), without taking into account the specimen to specimen differences in the size and spatial distribution of pores (Section 3.3.1, Figs. 3.2 and 3) that are likely related to the two difference processing techniques used in this study (HP and PECS). No experimental or theoretical studies in the open literature treat the porosity dependence of the elastic moduli of YbAl3 or any other trialuminide. Also, no experimental values of the elastic moduli for YbAl3 are available in the literature for any value of P. However, since E0, G0 and B0 (Eqs. (5a) – (5c)) represent estimates of the elastic moduli for P = 0, they can be compared directly to the theoretically calculated values of the elastic moduli (Table 3.6). The YbAl3 elastic modulus values determined experimentally in this study are presented in Table 3.6, along with a summary of calculated elastic modulus results for YbAl3 [Sa 2011; Zhou 2010; Tao 2008]. In addition, Table 3.6 includes calculated and experimental elastic modulus results for the additional rare earth trialuminide ScAl3 [Tao 2008; Hyland 1991; Fu 1990; Xu 1990; Jahnatek 2007] and TmAl3 [Tao 2008; Sa 2011; Zhou 2010]. Also, Table 3.6 includes a calculated bulk modulus value of a transition metal trialuminide, ZrAl3 [Xu 1990]. The three rare-earth and transition metal trialuminides have roughly similar elastic moduli, with E for the theoretical values ranging between 124.19 GPa and 165.36 GPa and the experimental YbAl3 value from this study 174.0 GPa. 68 Table 3.5. The fitting parameters obtained from the least-squares fit of the appropriate elastic modulus to Eqs. (5a) – (5c). The coefficient of determination, R2, is given for each of the fitted equations. Property Young's Modulus Shear Modulus Bulk Modulus P=0 Intercept (GPa) 173.8 ± 2.3 (E0) b (unitless) 2.35 ± 0.06 (bE) R2 73.5 ± 0.9 (G0) 2.38 ± 0.05 (bG) 0.995 91.6 ± 3.4 (B0) 2.21 ± 0.17 (bB) 0.947 0.994 Table 3.6. A comparison of this study’s experimentally determined values of elastic moduli (E, Young’s modulus, G, shear modulus, B, bulk modulus and , Poisson’s ratio) for YbAl3 with calculated and experimental values of other trialuminide materials, R-Al3, from the literature for R = Yb, Sc, Tm, or Zr [Sa 2011, Zhou 2010, Tao 2008, Hyland 1991, Fu 1990, Xu 1990, Jahnatek 2007]. Material E (GPa) YbAl3 174.0 ± 2.5 YbAl3 N.A. YbAl3 N.A. YbAl3 124.19 G (GPa) 73.6 ± 0.9 66.0 N.A. 54.17 B (GPa) 91.3 ± 2.9 N.A 67 58.51  0.182 ± 0.003 N.A N.A 0.146 Method Exp (RUS) Reference This study Calculated Calculated Calculated [Sa 2011] [Zhou 2010] [Tao 2008] ScAl3 ScAl3 69.62 68.4 ± 0.8 69 N.A. 78 88.24 91.5 ± 1.1 92 93 87 0.188 0.201 ± 0.001 0.20 N.A. N.A. Calculated Exp (Pulse echo) Calculated Calculated Calculated [Tao 2008] [Hyland 1991] ScAl3 ScAl3 ScAl3 165.36 164.2 ± 1.9 166 N.A. 174 TmAl3 TmAl3 TmAl3 ZrAl3 146.19 N.A. N.A. N.A. 61.48 60.3 N.A. N.A. 78.3 N.A 80 100 0.189 N.A N.A N.A. Calculated Calculated Calculated Calculated N.A. = Not available 69 [Fu 1990] [Xu 1990] [Jahnatek 2007] [Tao 2008] [Sa 2011] [Zhou 2010] [Xu 1990] 150 100 B0 50 0 0.0 G0 0.1 0.2 0.24 Poisson's Ratio Elastic Moduli (GPa) 200 Young's Modulus, PECS Shear Modulus, PECS Bulk Modulus, PECS Young's Modulus, HP Shear Modulus, HP Bulk Modulus, HP (a) E0 0.3 (b) Poisson's Ratio, PECS Poisson's Ratio, HP 0.22 0.20 0.18 0.16 0.0 0.1 0.2 0.3 P P Figure 3.5. For the YbAl3 specimens included in this study, the RUS measurement results for (a) the Young’s modulus, E, shear modulus, G, bulk modulus, B as a function of P and (b) Poisson’s ratio as a function of P. The solid lines in Fig. 3.5a represent leastsquares fits to Eqs. (5a), (5b) and (5c) for the Young’s modulus, shear modulus and bulk modulus, respectively. 70 Disc-shaped specimen YbAl3-C Tripod transducer arrangement (b) Bar-shaped specimen YbAl3-C Two transducer arrangement Intensity (AU) Intensity (AU) (a) 0 200 400 600 0 Frequency (kHz) Intensity (AU) (c) 0 200 400 600 Frequency (kHz) YbAl3-C (bar) Tripod transducer arrangement 200 400 600 Frequency (kHz) Figure 3.6. The RUS mechanical vibration spectra for the two-transducer stage and tripod specimen stage. 71 The values of the elastic moduli for YbAl3 theoretically predicted from DFT theory are considerably lower than the experimental modulus values obtained from this study. For example, the E0, G0 and B0 values of moduli theoretically predicted by Tao [Tao 2008] are approximately 29%, 21% and 36% lower, respectively, than the experimental modulus values obtained from this study (Table 3.6). Also the G0 value for YbAl3 calculated by Sa [Sa 2011] is about 10% lower than this study’s experimental value of G0 (Table 3.6) and Zhou’s calculated value for B0 [Zhou 2010] was about 27% lower than this study’s experimentally value of YbAl3 (Table 3.6). However, it is also critical to note that there are also significant disagreements among the theoretical elastic moduli obtained for YbAl3 (Table 3.6). In particular, the G0 values calculated by Tao [Tao 2008] and Sa [Sa 2011] differ by about 18% (Table 3.6). Also, the B0 values calculated by Zhou [Zhou 2010] and Tao [Tao 2008] differ by approximately 13% (Table 3.6). Thus, although each of these three theoretical studies are relatively recent (2008, 2010 and 2011) for the Tao et al. [Tao 2008], Zhou et al. [Zhou 2010] and Sao [Sao 2011] studies, respectively, the differences among these theoretical studies of the elastic moduli YbAl3 indicate a difficulty in obtaining consistent theoretical predictions. In contrast to the lack of agreement among the theoretical studies/experimental studies for the elastic moduli of YbAl3, the situation appears to be different for another rare earth trialuminide, namely ScAl3 where theoretical values for B0 for four theoretical studies [Tao 2008; Fu 1990; Xu 1990; Jahnateck 2007] are each within about 6 - 10% of one another (Table 3.6). (B0 is used here as a basis of comparison for the theoretical studies, as shown in Table 3.6, since B0 is the only elastic modulus that is common to the four theoretical studies as well as the experimental study). The 72 experimentally determined moduli E0, G0 and B0 [Hyland 1991] for ScAl3 all lie within the range defined by the maximum and the minimum of the theoretical estimates [Tao 2008, Fu 1990, Xu 1990, Jahnateck 2007] (Table 3.6). In addition, although the agreement may be fortuitous, it is interesting to note that each of the E0, G0 and B0 values calculated by Jahnateck et al. [Jahnateck 2007] for ScAl3 agree relatively well with the experimental values obtained for YbAl3 in this study (Table 3.6). Furthermore, for yet another rare earth trialuminide, TmAl3, the B0 and G0 values calculated among the theoretical studies differ by roughly 2% (Table 3.6) [Tao 2008, Sa 2011, Zhou 2011. However, the experimental values for the elastic moduli of TmAl3 are not available in the literature to allow comparison among the various theoretical predictions and experimental data. Thus, among the three rare earth trialuminides, YbAl3, ScAl3 and TmAl3, the elasticity results summarized in Table 3.6 indicate that the difficulty in obtaining consistent results for the elastic moduli is much more pronounced for YbAl3 than for ScAl3 and TmAl3. The apparent difficulty in calculating the elastic moduli for YbAl3 in particular may be related to complexities in modeling and theoretically understanding the intermediate valent nature of the Yb ion in these compounds [4] and its effect on the elastic properties. As an internal check on the RUS data included in our study, the elastic moduli for selected YbAl3 disc-shaped specimens were measured on the tripod transducer stage for the RUS (Fig. 3.1a).The specimen was then cut into a bar shaped specimen and measured using the two-transducer stage (Fig. 3.1b). The significance of these measurements with the differing specimen geometries (disk and bar) and differing modes of excitation (tripod 73 and two-transducer or bipod stages) is that mechanical resonant mode patterns are sensitive on specimen geometry [Rossing 1982, Ogi 2002] although the mechanical resonances excited in the two-transducer stage should be similar except for the relative intensities. The differences in the resulting RUS resonant spectra are directly demonstrated in Fig. 3.6, which shows differing resonance frequencies and relative intensities for the bar specimen YAl3-C on the tripod transducer stage (Fig. 3.6a), the disc geometry on the transducer tripod (Fig. 3.6b) and the bar specimen on the two transducer stage. Specimen YAl3-C specimen was first measured as a disc on the tripod transducer stage, then it was cut into a bar and measured on the both the tripod and two transducer stages. Although the RUS measurements for the specimen YAl3-C was thus determined in three separate experiments, the E and G values measured for the various transducer stages/specimen geometries (Fig. 3.5) differ by less than 1 percent (Table 3.7), indicating a great deal of internal consistency in the measurements. As an example of modulus-porosity study for a thermoelectric material other than YbAl3, for 12 specimens of the thermoelectric LAST (lead-antimony-silvertellurium) with the compositions Ag0.43Pb18Sb1.2Te20 and Ag0.86Pb19Sb1.0Te20 having volume fraction porosities, P, ranging from 0.01 to 0.14, the Young’s modulus and shear modulus versus P behavior was described well by Eqs. (5a) and (5b) respectively [Ni 2009]. In particular, for the Young’s modulus versus P data for LAST, a least-squares fit to Eq. (5a) yielded E0 = 58.3 ± 0.3 GPa, bE = 3.6 ± 0.1 with a correlation coefficient, r2 = 0.994 [Ni 2009]. Thus, Eq. (5a) describes the Young’s modulus versus porosity behavior relatively well for both LAST and YbAl3. 74 Table 3.7. For the seven YbAl3 specimens included in this study, N, the number of RUS resonant peaks measured, the RMS error in RUS, the Young’s modulus, E, shear modulus, G, Poisson’s ratio, ν, longitudinal and shear velocity, VL and VS, acoustic Debye temperature, θD, and porosity, P, were measured. The RUS analysis on each specimen was performed on the tripod configuration (Fig. 3.1a). Specimen Label N RMS Error E (GPa) G (GPa) ν VL (m/s) VS (m/s) θD (K) P YbAl3-A (bar) 20 0.19% 126.2 ± 0.3 51.65 ± 0.05 0.222 ± 0.001 5390 3220 384 0.125 YbAl3-B (disc) 39 0.41% 77.1 ± 0.2 32.08 ± 0.05 0.201 ± 0.001 4440 2710 308 0.233 YbAl3-B (bar) 24 0.27% 85.8 ± 0.3 36.19 ± 0.05 0.186 ± 0.002 4590 2850 326 0.217 YbAl3-C (disc) 28 0.21% 152.1 ± 0.2 64.42 ± 0.05 0.180 ± 0.001 5560 3470 421 0.060 YbAl3-C (bar, tripod) 17 0.25% 152.3 ± 0.4 64.51 ± 0.08 0.181± 0.002 5550 3460 421 0.054 YbAl3-C (bar, two transducer) 19 0.33% 152.9 ± 0.5 64.89 ± 0.10 0.178 ± 0.002 5550 3470 422 0.054 YbAl3-D (bar) 25 0.28% 156.7 ± 0.5 66.37 ± 0.08 0.181 ± 0.002 5560 3470 426 0.030 YbAl3-E (bar) 22 0.21% 79.7 ± 0.2 33.2 ± 0.04 0.202 ± 0.002 4490 2750 313 0.227 75 Over the range from P = 0.030 to 0.060, the Poisson’s ratio, ν, of YbAl3 is relatively constant at ν = 0.18 (Fig. 3.5b). For P = 0.125 the Poisson’s ratio increases 0.21, then as P increases further ν ranges from about 0.19 to 0.20 (Fig. 3.5b). In the literature for oxide materials, Boccaccini found empirically that, although there was considerable scatter in the ν versus P data, the increase or decrease in ν with increasing P seemed to be correlated with the value of ν0 (ν at P = 0), that is, where ν increased with increasing P for ν0 > 0.25 and ν decreased with increasing P for ν0 < 0.25 [Boccaccini 1994]. In this study, the apparent scatter in ν between values of about 0.18 to 0.21 makes it difficult to discern a trend in the P dependence of ν for YbAl3 (Fig. 3.5b). 3.3.5. Acoustic wave speeds and the Debye temperature The longitudinal acoustic wave speed, VL and the shear acoustic wave speed, VS, are also obtained from the RUS analysis. For specimens with low to intermediate values of P, the acoustic velocities can be approximated as a linear function of P [Aliev 2011, Solunke 2007] such that VL  VL0 1  cL P  (6a) VS  VS 0 1  cS P  (6b) where VL0 and VS0 are the acoustic wave speeds corresponding to a theoretically dense material and cL and cS are material-dependent constants. For the YbAl3 specimens included in this study, Eqs. (6a) and (6b) describe relatively well the P dependence of the acoustic wave speeds (Table 3.8, Fig. 3.7a). 76 Table 3.8. The fitting parameters obtained from the least-squares fit of the VL and VS versus P data to Eqs. (6a) and (6b), respectively, with the coefficient of determination, R2. Equation P=0 Intercept (mm/µs) 5.88 ± 0.09 (vl0) 3.67 ± 0.04 (vs0) Acoustic Velocity (m/s) Longitudinal (6a) Velocity Shear (6b) Velocity Longitudinal Velocity, PECS Shear Velocity, PECS Longitudinal Velocity, HP Shear Velocity, HP (a) 6 5 vL0 4 vS0 3 0.0 0.1 0.2 0.3 c (unitless) R2 1.00 ± 0.09 0.958 1.07 ± 0.06 0.986 Debye Temperature (K) Property P (b) Debye Temperature, PECS Debye Temperature, HP 450 400 D0 350 300 0.0 0.1 0.2 0.3 P Figure 3.7. From the RUS measurements of the YbAl3 specimens included in this study, (a) shear and longitudinal wave speeds as a function of P, where the solid lines in represent least-squares fits to Eq. (6a) and (6b) for the longitudinal and shear velocities, respectively. (b) the effective Debye temperature as a function of P, where the solid line represents a least-squares fit to Eq. (9). 77 The Anderson approximation [Anderson 1963] was used to calculate θD, the acoustic Debye temperature h D  kB  3q N A    4 M  Vm   1/ 3 (7) where h is Planck’s constant, kB is Boltzmann’s constant, NA is Avogadro’s number, ρ is density, M is molecular weight, and q is the number of atoms per molecule. The average acoustic velocity, VM, was calculated from VL and VS using the relationship [Anderson 1963] 1 2 1  VM    3  3    3 VS VL   1 / 3 (8) The linear decrease in the effective Debye temperature with increasing porosity (Fig. 3.7b) and can be described by  D   D 0 1  b P  (9) where a least-squares fit of the θD versus P data gives fitting parameters θD0 = 453 + K, bθ = 1.32 + 0.06 with a coefficient of determination of 0.99. 3.3.6 Mechanical properties as a function of processing technique In this study, the YbAl3 specimens were densified either by HP or PECS processing (Table 3.2). Although some researchers have reported that differing processing techniques can result in differing mechanical properties [Rice 1998], the Vickers indentation hardness measurements (Fig. 3.4) and the RUS elasticity measurements for the Young’s, shear and bulk moduli (Fig. 3.5a) as well as the acoustic velocity (Fig. 3.7a) and the Debye temperature (Fig. 3.7b) show no significant differences 78 in the magnitudes of H, E, G, B, VL, VS or θD as a function of two processing techniques (HP or PECS). Thus, although the HP and PECS processing parameters (sintering temperature and sintering time) yield differing sizes and spatial distribution of pores (Figs. 3.2 and 3.3 and Section 3.3.1), the physical properties measured for the YbAl3 specimens included in this study were functions of the volume fraction porosity P and not the processing details. The practical significance of this is that for the two most common techniques of densifying thermoelectric materials, namely HP and PECS, the property-porosity relationships found in this study for YbAl3 do not appear to be functions of the processing technique. 3.4 Summary and Conclusions For the low temperature thermoelectric material YbAl3, the room temperature hardness, Young’s modulus, shear modulus, bulk modulus and Poisson’s ratio were measured using RUS for polycrystalline specimens with volume fraction porosities, P, over the interval 0.030 < P < 0.233. The rates of linear decrease in hardness with increasing P (Fig. 3.4, Eq. (4b)) and the elastic moduli (Fig. 5, Eqs. (5a) – (5c)) are consistent with the literature. Although no fracture toughness literature is available for rare earth trialumindes, the YbAl3 values of approximately 1.1 to 1.6 MPa m1/2 that were found in this study are comparable to lower ranges of KC measured for transition metal trialumindes, for which significant efforts have been directed toward enhancing KC for structural materials applications. Therefore, it appears that the KC of YbAl3 may be sufficiently high so as not 79 to hinder its development as a low temperature thermoelectric material. Also, the hardness value 6.81 GPa for YbAl3 (P = 0.030) is comparable to the literature hardness value of the transition metal trialuminide TiAl3, thus the hardness of YbAl3 is likely sufficient to impart reasonable in-service scratch and abrasion resistance. However, once again, the lack of hardness data in the rare earth trialuminide literature limits the hardness comparisons that can be made. For this study, extrapolation of the elastic moduli versus porosity results to zero porosity gives intercepts E0, G0 and B0 for YbAl3 (Eqs. (5a) – (5c), Fig. 3.6). These zero porosity values are roughly 10 - 36% higher than the E0, G0 and B0 values predicted from density functional theory (Table 3.6). However, the E0, G0 and B0 values predicted theoretically for YbAl3 are also significantly different from one another (Table 3.6). In contrast, the theoretical and experimental moduli for ScAl3 are relatively successful. The apparent difficulty in calculating the elastic moduli for YbAl3 may be related to complexities in modeling and understanding the nature of the intermediate valence nature of the Yb ion in YbAl3. ACKNOWLEDGEMENTS GJL and DTM. acknowledge support by the Air Force Office of Scientific Research under the MURI program “Cryogenic Peltier Cooling,” Contract #FA9550-101-0533. In addition, R.D.S. and E.D.C. acknowledge the financial support of the Department of Energy, “Revolutionary Materials for Solid State Energy Conversion Center,” an Energy Frontiers Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic energy Sciences under award number DE-SC0001054. 80 REFERENCES 81 REFERENCES [Aliev 2011] G. N. Aliev, B. Goller, and P. A. Snow, Elastic properties of porous silicon studied by acoustic transmission spectroscopy, J. Appl. Phys. 110, (2011).043534 [Anstis 1981] GR Anstis, P Chantikul, BR Lawn, DB Marshall, A Critical Evaluation of Indentation Techniques for Measuring Fracture Toughness: I, Direct Crack Measurements Journal of the American Ceramic Society, 64 (1981) 533 – 538. [Biswas 2011] K. Biswas, J. He, Q. Zhang, G. Wang, C. Uher, V. P. Dravid, and M. G. Kanatzidis, “Strained endotaxial nanostructures with high thermoelectric figure of merit, Nature Chemistry 3 (2011) 160 [Boccaccini 1994] A.R. Boccaccini, J. Am. Ceram. Soc. 77 (1994) 2779. [Canfield 1992] P.C. Canfield, Z. Fisk, Growth of single crystals from metallic fluxes, Philosophical Magazine B 65, 1117 (1992). [Fu 1990] C. L. Fu, Electronic, elastic, and fracture properties of trialuminide alloys: Al3Sc and Al3Ti, J. Mater. Res. 5 (1990)971 – 979. [Havinga 1973] E.E. Havinga, K.H.J. Buschow and H.J. van Daal, The ambivalence of Yb in YbA12 and YbA13, Solid State Communications 13 (1973) 621. [Hu 2011] C. Hu, Y. Sakka1, Toshiyuki Nishimura, S. Guo, S. Grasso. H, Tanaka, Physical and mechanical properties of highly textured polycrystalline Nb4AlC3 ceramic, Sci. Technol. Adv. Mater. 12 (2011) 044603. [Hyland 1991] R. W. Hyland, R. C. Stiffler, Determination of the elastic constants of polycrystalline of Al3Sc, Scripta Metallurgica 25 (1991) 473 – 477. [Jahnatek 2007] M. Jahnatek , M. Krajcí , J. Hafner, Interatomic bonds and the tensile anisotropy of trialuminides in the elastic limit: a density functional study for Al3(Sc, Ti, V, Cr), Philosophical Magazine 87 (2007) 1769 – 1794. [Jang 2004] H.S. Jang, C.W. Kang, Y. Kim, K.T. Hong, S.-J. Kim, Effects of Mn additions on microstructure and mechanical properties of (Al + x at.% Mn)3Ti intermetallic compounds prepared by mechanical alloying and spark plasma sintering, Intermetallics 12 (2004) 477 – 485. [Lee 2006] S.H. Lee, K.I. Moon, K.S. Lee, Enhancement of the fracture toughness of bulk L12-based (Al + 12.5 at.% M)3Zr (M = Cu, Mn) intermetallics synthesized by mechanical alloying, Intermetallics 14 (2006) 1 – 8. 82 [Mahan 1998] G.D. Mahan, in H. Ehrenreich and F. Spaepen, editors, “Good Thermoelectrics,” in Solid State Physics: Advances in Research and Applications, Academic Press, San Diego, CA, volume 51, 1998 page 137. [Massalski 1987] T.B. Massalski editor, Binary Alloy Phase Diagrams, ASM International, Materials Park, OH, volume 1, 1987 page 238. [Migliori 1997] A Migliori, JL Sarrao, Resonant ultrasound spectroscopy applications to physics, materials measurements and nondestructive evaluation, WileyInterscience, Hoboken, NJ, USA, 1997. [Milman 2001] Y. V. Milman, D.B. Miracle, S.I. Chugunova, I.V. Voskoboinik, N.P. Korzhova, T.N. Legkaya, Y.N. Podrezov, Mechanical behaviour of Al3Ti intermetallics , Intermetallics 9 (2001) 839–845 [Milman 2011] Y.V. Milman, A.A. Golubenko, S.N. Dub, Indentation size effect in nanohardness, Acta Materialia 59 (2011) 7480–7487 [Ni 2009] J. E. Ni, F. Ren, E. D. Case, E. J. Timm, Porosity dependence of elastic moduli in LAST (Lead-antimony-silver-tellurium) thermoelectric materials, Materials Chemistry and Physics 118[2-3] (2009) 459 – 466. [Ni 2010] J. E. Ni, E. D. Case, K. Khabir, C.-I Wu, T. P. Hogan, E. J. Timm, S. Girard, M. G. Kanatzidis, Room temperature Young’s modulus, shear modulus, Poisson’s ratio and hardness of PbTe-PbS thermoelectric materials, Materials Science and Engineering B 170 (2010) 58 – 66. [Ni 2012] J. E. Ni, E. D. Case, R. Stewart, C.-I Wu, T. Hogan, M. G. Kanatzidis, , Bloating in (Pb0.95Sn0.05Te)0.92(PbS)0.08-0.055% PbI2 specimens as a result of processing conditions, Journal of Electronic Materials, 41 (2012) 1153 – 1158. [Nix 1998] W.D. Nix, H. Gao, J. Mech. Phys. Solids 46 (1998) 411 – 425. [Ogi 2002] H. Ogi, K. Sato, T. Asada , M Hirao Complete mode identification for resonance ultrasound spectroscopy, J. Acoust. Soc. Am. 112 (2002) 2553 – 2557. [Ran 2010a] S. Ran, L. Zhang, O. Van der Biest, J. Vleugels, Pulsed electric current, in situ synthesis and sintering of textured TiB2 ceramics, Journal of the European Ceramic Society 30 (2010) 1043 – 1047. [Ran 2010b] S. Ran, O. Van der Biest, J. Vleugels, ZrB2–SiC composites prepared by reactive pulsed electric current sintering, Journal of the European Ceramic Society 30 (2010) 2633 – 2642. 83 [Ren 2008] F. Ren, E. D. Case, J. R. Sootsman and M. G. Kanatzidis, H. Kong, C. Uher E. Lara-Curzio and R. M. Trejo, C. Uher, The high temperature elastic moduli of polycrystalline PbTe measured by resonant ultrasound spectroscopy, Acta Materialia 56 (2008) 5954 – 5963. [Ren 2009a] F. Ren, E. D. Case, J. E. Ni, E. J. Timm, E. Lara-Curzio, R. M. Trejo, C.-H. Lin, M. G. Kanatzidis, 2009, Temperature-dependent elastic moduli of leadtelluride based thermoelectric materials, Philosophical Magazine, 89 (2009) 143 – 167. [Rice 1998] R. W. Rice, Porosity of Ceramics, Marcel-Dekker, Inc., New York. 1998. [Rossing 1982] T.D. Rossing, Chladni’s law for vibrating plates, Am. J. Phys. 50 (1982) 271 – 274. [Rowe 2002] D.M. Rowe, V.L. Kuznetsov, L.A. Kuznetsova, and G. Min, Electrical and thermal transport properties of intermediate-valent YbAl3, J. Phys. D: Applied Physics 35 (2002) 2183. [Sa 2011] B. Sa, J. Zhou, Z. Sun, Electronic mechanism of shear modulus enhancement in rare earth intermetallics Yb1_xTmxAl3, Intermetallics 19 (2011) 1020 – 1023. [Sangwal 2000] K. Sangwal, Mater. Chem. Phys. 63 (2000) 145 – 152. [Schmidt 2012] R. D. Schmidt, E. Case, J.E. Ni, J. Giles III, T. P. Hogan, Roomtemperature mechanical properties and slow crack growth behavior of Mg2Si thermoelectric materials, Journal of Electronic Materials, 41 (2012) 1210 – 1216. [Shi 2011] X. Shi, J. Yang, J. R. Salvador, M. Chi, J. Y. Cho, H. Wang, S. Bai, J. Yang, W. Zhang, and L. Chen, “Multiple-filled skutterudites: high thermoelectric figure of merit through separately optimizing electrical and thermal transports,” Journal of the American Chemical Society 133,7837 (2011). [Solunke 2007] M.B. Solunke, P.U. Sharma, V.K. Lakhani, M.P. Pandya, K.B. Modi, P.V. Reddy, S.S. Shah, Elastic properties of gallium substituted Bi (Pb)-2212 superconducting system at 300 K, Ceramics International 33 (2007) 21 – 26. [Tao 2008] X. Tao, Y. Ouyang, H. Liu, Y. Feng, Y. DU, Z. Jin, First Principles Calculations of the Thermodynamic and elastic properties of the L12-based Al3RE (RE = Sc, Y, La – Lu), International Journal of Materials Research, 99 (2008) 582 – 588. [Van Daal 1974] H.J. Van Daal, P.B. Van Aken, and K.H.J Buschow, The Seebeck coefficient of YbAl2 and YbAl3, Physics Letters 49A (1974) 246. 84 [Varin 2001] R.A. Varin, L. Zbroniec, T. Czujko, Y.-K. Song, Fracture toughness of intermetallic compacts consolidated from nanocrystalline powders, Mater. Sci. Eng A300 (2001) 1 – 11. [Wachtman 2009] JB Wachtman, WR Cannon, Matthewson MJ, Mechanical Properties of Ceramics, Second Edition, John Wiley & Sons, Inc., New York, 2009. [Weaver 2003] M.L. Weaver, M.E. Stevenson, R.C. Bradt, Mater. Sci. Eng. A345 (2003) 113 – 117. [Xu 1990] J.-H. Xu, A.J. Freeman, Phase stability and electronic structure of ScAl3 and ZrAl3 and of Sc-stabilized cubic ZrAl3 precipitates, Phys. Rev. B 41 (1990) 12553 – 12561. [Zhou 2010] J Zhou, BS Sa, ZM Sun., First-principles investigations on phase stability and electronic structures of Yb(1-x)M(x)Al(3) (M = Ho, Er and Tm) alloys, Intermetallics 18 (2010) 2394 – 2399. 85 4 Room-Temperature Mechanical Properties and Slow Crack Growth Behavior of Mg2Si Thermoelectric Materials Robert D. Schmidt1, Eldon D. Case1, Jesse Giles III1, Jennifer E. Ni1, Timothy P. Hogan2 1 Chemical Engineering and Materials Science Department, Michigan State University, East Lansing, MI, 48824 2 Electrical and Computer Engineering Department, Michigan State University, East Lansing, MI, 48824 Published in: Journal of Electronic Materials 41 (2012) 1210 – 1216. Abstract Mg2Si is of interest as a thermoelectric (TE) material in part due to its low materials cost, lack of toxic components, and low mass density. However, harvesting of waste heat subjects TE materials to a range of mechanical and thermal stresses. To understand and model the material’s response to such stresses, the mechanical properties of the TE material must be known. The Mg2Si specimens included in this study were powder processed and then sintered via pulsed electrical current sintering. The elastic moduli (Young’s modulus, shear modulus, and Poisson’s ratio) were measured using resonant ultrasound spectroscopy, while the hardness and fracture toughness were examined using Vickers indentation. Also, the Vickers indentation crack lengths were measured as a function of time in room air to determine the susceptibility of Mg2Si to slow crack growth. Keywords: Magnesium silicide, fracture toughness, elastic modulus, hardness, slow crack growth 86 4.1 Introduction New thermoelectric (TE) materials based on Mg2Si are candidates for a wide variety of waste heat recovery applications due to their low mass density [~2 g/cm3 for Mg2Si versus densities of 6 g/cm3 to 8 g/cm3 for TE materials such as PbTe, leadantimony-silver-tellurium (LAST), CoSb3, FeSb3,and Bi2Te3]. Also, Mg2Si is composed of abundant, inexpensive, and nontoxic elements. Previous studies have determined a ZT as high as 1.0 at 800 K for an Mg2Si-based material [Mars 2009]. Energy harvesting applications subject TE materials to thermal cycling during the startup and operation of the device, as well as to the thermal gradient required to operate a TE generator. The designer must know the mechanical properties such as the elastic moduli in order to calculate the material’s response to imposed thermal and mechanical loads. The roomtemperature elastic moduli, hardness, and fracture toughness of Mg2Si are determined in this study. In addition, since fine powders of Mg2Si react vigorously with water, [Aesar] a slow crack growth study was performed in which the length of Vickers radial indentation cracks was monitored as a function of time in room air. Slow crack growth involves a concerted chemical reaction [Michalsky 1983] of a polar molecule such as water with the highly strained atomic bonds at a crack tip, with the result that the reaction breaks bonds at the crack tip and the crack slowly advances [Wiederhorn 1968; West 1998; Freiman 2009]. Slow crack growth has been observed in a variety of brittle materials, including ceramics [Freiman 2009] as well as in single crystal silicon [Connally 1992]. If slow crack growth occurs in Mg2Si, this would imply that the flaws produced by cutting and grinding procedures during module preparation could extend upon exposure to either room air or an aqueous environment. 87 Temperature Pressure 50 1000 40 800 30 600 20 400 10 200 0 5 10 15 20 Pressure (MPa) Temperature (K) 1200 0 25 Time (min) Figure 4.1. The time–temperature and time–pressure profiles used in the PECS processing of each of the Mg2Si specimens included in this study. Table 4.1. The Mg2Si specimens used in this study for Vicker’s hardness, H, fracture toughness, KIC, Young’s modulus, E, shear modulus, G, Poisson’s ratio, , and acoustic debye temperature, D, measurements. The Mg2Si pieces were reduced using a (1) mechanical mortar and pestle until all powder passed through a 53 µm sieve and then either densified or further reduced using a (2) planetary mill at 150 rpm for 3 hours. For indentation the specimens were polished to 1 μm diamond grit, further details are available elsewhere [Ni 2010]. Due to reactions with water [Aesar], the specimens were cleaned using ethanol instead of water between polishing grit sizes. Specimen Powder Processing Form Name Dimensions Mass Density (mm) (g) (g/cm3) Mechanical properties measured Mg2Si-01 Planetary mill Disc 19.55 dia 3.35 ht 2.01 2.01 H, KICa, b Mg2Si-03 Planetary mill Parallele- 9.89 x 13.01 0.90 piped x 3.35 2.09 E, G, , Dc Mg2Si-05 Mechanical mortar and pestle Disc 12.54 dia 3.41 ht 0.88 2.08 H, KICa Mg2Si-11 Mechanical mortar and pestle Parallele- 11.45 x 1.18 piped 14.86 x 3.31 2.09 E, G, , Dc Mechanical mortar Parallele- 11.36 x 1.07 2.07 E, G, , Dc and pestle piped 15.06 x 3.01 a Measured using Vicker’s indentations, b Slow crack growth monitored using this specimen, c Measured using RUS 88 Mg2Si-12 4.2 Experimental procedure Mg2Si (99.99% metals basis purity), purchased from Alfa Aesar (45518, Ward Hill, MA), was reduced in size using a mechanical mortar and pestle and planetary milling (Table 4.1). Further details of the powder processing are presented elsewhere [Ni 2010]. The powders were densified using pulsed electronic sintering (PECS) at 1073 K for 5 min and 50 MPa pressure (Fig. 4.1) in argon. The grain size was measured using the linear intercept technique with a stereographic projection factor of 1.5 [Underwood 1968; Case 1981]. All of the grain size analysis on the Mg2Si specimens was performed on micrographs of fractured surfaces taken in the secondary electron mode on a scanning electron microscope (SEM). The elastic moduli (Table 4.1) were calculated using a dynamic method, resonant ultrasound spectroscopy (RUS), and commercially available software. Additional details of the RUS elasticity technique are provided elsewhere [Ni 2010; Ren 2009]. The Vickers hardness, H, and fracture toughness, KIC, were determined using Vickers indentations with loads ranging from 2.94 N to 9.81 N. H was calculated by [Wachtman 2009] ( (1) ) where P is the indenter load and 2a is the length of the indentation diagonal. KIC was estimated from [Wachtman 2009] ( ⁄ ) ⁄ (2) ⁄ 89 where E is the Young’s modulus calculated using RUS, c is the radial crack length extending from the center of a Vickers indentation, and ξ is a dimensionless constant assumed to be 0.016 [Wachtman 2009]. The Vickers indentation was performed by a Buehler indenter (Buehler Semimacro Indenter, Lake Bluff, IL) and a Shimadzu hardness tester (Shimadzu HMV2000, Kyoto, Japan). Each indenter was calibrated using a steel calibration block with H of 7.75 GPa (761-048; Yamamoto Scientific Tools Lab, Co LTD, Japan). For both the hardness calibration and for each Vickers indentation performed on the Mg2Si specimens in this study, a load time of 5 s and a loading speed of 70 µm/s were used. Also, both the calibration block and the Mg2Si specimens were indented at loads of 2.95 N, 4.90 N, and 9.81 N, with 10 indentations performed for each load on a given specimen. The radial crack length and the diagonal length of the indentation impression were measured optically using the movable filars on the Buehler and Shimadzu indentation machines. To compare the hardness results obtained from the Buehler and Shimadzu indentation machines, a Mg2Si specimen was indented 10 times at each load of 2.95 N and 4.90 N by each machine at a load time of 5 s and a loading speed of 70 µm/s. Possible slow crack growth was monitored using radial cracks induced by Vickers indentation at 2.94 N load. The radial cracks were measured at intervals between 1 min and 3 days after indentation. During the slow crack growth study, the room temperature and relative humidity ranged between 23.2°C and 24.3°C and 46% and 51%, respectively. 90 Counts (Arbitrary units) 400 MgO Mg2Si 300 200 100 0 20 30 40 50 60 70 2 Figure 4.2. X-ray diffraction of planetary-milled Mg2Si powder with peaks compared with the literature for MgO [Hazen 1976] and Mg2Si [Owen 1923]. Figure 4.3. Fractured surface of specimens (a) Mg2Si-01 and (b) Mg2Si-05, PECS processed from planetary mill and mechanical mortar and pestle Mg2Si powder, respectively. Cleavage steps show mixed- mode fracture, with both intergranular and transgranular fracture present. Entrapped pores visible at grain boundaries were approximately 1 µm to 3 µm diameter with polygonal shape. The dark areas between grains are porosity within the specimens, typically around 1 µm or less in diameter 91 Figure 4.4. Typical Vickers indentation crack systems for specimens Mg2Si-01 and Mg2Si-05, which display indentation impressions, little chipping or spalling, and a fully developed radial crack system. Table 4.2. The density, Young’s modulus, E, shear modulus, G, Poisson’s ratio, , and Debye temperature, D, from this study compared with the literature. The moduli were determined using resonant ultrasound spectroscopy, RUS, first principle calculation, FPC, resonance technique, RT, and compression. Reference Modulus Density Measurement E(GPa) (g/cm3) Method This Study (Mg2Si-03) 2.09 RUS This Study (Mg2Si-11) 2.09 This Study (Mg2Si-12) G (GPa)  D (K) 117.2 ± 0.2 49.53 ± 0.04 0.183 ± 0.001 585a RUS 117.7 ± 0.2 50.09 ± 0.05 0.175 ± 0.001 588a 2.07 RUS 116.9 ± 0.5 48.92 ± 0.13 0.195 ± 0.001 583a [Zhang 2009] 1.98b FPC 115.6 49.50 0.17 N/A [Tani 2008] 2.04b FPC 113.5 49.2 0.161 432 RT 110.9c 47.6c 0.165c 578 [Whitten 1965] 2.00 d [Milekhine 1.94e Compression 76 ± 6 N/A N/A N/A 2002] N/A Not available, a Acoustic Debye temperature, b Density calculated from the given lattice parameter, c From the average of Hashin and Shtrikman bounds at 300 K [Simmons 1971], d Debye temperature at 0 K, e Density estimated by 3% porosity and 2.00 g/cc porosity free density 92 4.3 Results and Discussion In this study, the Mg2Si specimens had a mean mass density of 2.07 g/cm3, calculated from the measured mass and dimensions (Table 4.1). Single- crystal Mg2Si has mass density of 1.98 g/cm3 and 2.00 g/cm3 [Zhang 2009; Whitten 1965]. X-ray diffraction determined that an MgO phase [Hazen 1976] was present in the Mg2Si powder [Owen 1923] used to fabricate the specimens (Fig. 4.2), which may account for the increase in density for the PECS Mg2Si specimens in this study. The grain size was 2.4 µm (Fig. 4.3a) and 3.9 µm (Fig. 4.3b) for the specimens fabricated using powder reduced by the planetary mill and mortar and pestle, respectively. Based on SEM micrographs of the specimens (Figs. 4.3, 4.4), we estimate volume fraction porosity to be 0.05 or less, and a resulting approximate volume fraction of 0.89 ± 0.03 Mg2Si and 0.08 ± 0.03 MgO to account for the measured mass density. In this study, the mean values of the measured elastic moduli for Mg2Si were 117.3 GPa for Young’s modulus, E, 49.5 GPa for shear modulus, G, and 0.184 for Poisson’s ratio, ν. In the literature, E ranged from 110.9 GPa to 115.6 GPa,[Zhang 2009; Whitten 1965; Simmons 1971; Tani 2008] which was comparable to this study’s mean E of 117.3 GPa. The exception was Milekhine et al., who reported an E of 76 GPa; however, Milekhine et al. recognized that their Young’s modulus value was not consistent with literature [Milekhine 2002] (Table 4.2). For comparison with other TE materials, E values include CoSb3 at 140.6 GPa [Schmidt 2010] and PbTe-PbS 8% at 53.1 GPa [Ni 2010]. The calibration factors from 10 indentations at each load for the Shimadzu were 0.98, 0.96, and 0.98 at 2.95 N, 4.90 N, and 9.81 N loads, respectively, and for the Buehler 93 were 1.10, 1.10, and 1.03 at the same loads. The calibration factor was multiplied by the uncorrected H value for the reported H values in this study. For indentation loads of 2.95 N, 4.90 N, and 9.81 N, the Vickers hardness, H, in this study ranged from 4.8 ± 0.3 GPa to 5.6 ± 0.2 GPa. Over this relatively restricted load range, no load dependence of H was observed. Also, based on the mean and standard deviation of the measured H using the Shimadzu and Buehler indenters, there was no statistically significant difference in the H values measured, although the measured H was different from literature values [Milekhine 2002; Li 1993] (Table 4.3). The average hardness of 5.1 GPa for the PECS-processed specimens in this study was higher than the range of 3.96 GPa to 4.38 GPa for cast specimens in the literature [Milekhine 2002; Li 1993] (Table 4.3). In general, hardness increases with decreasing grain size [Ni 2010; Lawn 1993; Barsoum 2003]. The grain sizes of the cast speci- mens included in their studies were not given by Milekhine et al. [Milekhine 2002] and Li et al. [Li 1993]. However, grain sizes for cast specimens are often greater than 500 µm [Ni 2010; Ren 2008], which is significantly larger than the roughly 33 µm or smaller Vickers impression dimensions reported by Milekhine et al. [Milekhine 2002]. The likely grain size for the cast specimens is also much larger than the average grain sizes of 2.4 µm for the Mg2Si-01 and 3.9 µm for the Mg2Si-05 specimen included in this study. The narrow range of grain sizes (2.4 µm to 3.9 µm) makes it extremely unlikely that an H dependence on grain size would be observed for the specimens included in this study. However, the H value of 5.1 GPa for this study is expected to be higher than the H values of 3.96 GPa to 4.38 GPa measured for cast specimens [Milekhine 2002; Li 1993], since the cast specimens likely 94 Table 4.3. A comparisons of Vickers hardness, H, and fracture toughness, KIC, from this study to the literature. For the PECS processed specimens in this study, the H and KIC was determined using at least 10 indentations per load with one standard deviation reported as error. For cast specimens in literature, Milekhine et al. measured at least 10 indentations per load [Milekhine 2002]and Li et al. measured 5 indentations [Li 1993]. Reference Grain Size (µm) This Study (Mg2Si-01) 2.4 Load 0.981 N H (GPa) This Study (Mg2Si-05) 3.9 [Milekhine 2002] N/A [Li 1993] Mechanical Property N/A KIC (MPa·m1/2) H (GPa) KIC (MPa·m1/2) H (GPa) KIC (MPa·m1/2) N/A 2.94 N 4.90 N 9.81 N 5.4 ± 0.2a 5.6 ± 0.2a a b b 5.4 ± 0.2 5.0 ± 0.3 5.0 ± 0.1 1.2 ± 0.2b 1.3 ± 0.3b 1.3 ± 0.5b 4.8 ± 0.2b 4.8 ± 0.2b 5.0 ± 0.2a N/A N/A H (GPa) 4.38 ± 0.07 KIC (MPa·m1/2) N/A a Using a Buehler indenter Using a Shimadzu indenter N/A = not available b 95 0.7 ± 0.2b 1.1 ± 0.2b 0.9 ± 0.1b 3.96 4.05 4.20 0.88 0.81 0.74 N/A N/A N/A have much larger grain sizes. For other TE materials, the Vickers hardness is 1.10 GPa to 1.28 GPa for PbTe-PbS8% when powder processed [Ni 2010], 0.51 GPa to 1.2 GPa for lead-antimony- silver-tellurium (LAST) [Ren 2008], and 5.7 GPa for MM0.7Fe3CoSb12 (where MM denotes mischmetal) [Zhang 2010]. Fracture toughness, KIC, by Vickers indentation (Fig. 4.4) was 0.7 ± 0.2 MPa m1/2 to 1.3 ± 0.5 MPa m1/2 for indentation loads from 2.94 N to 9.81 N (Table 4.3). Milekhine et al. reported KIC values ranging from 0.81 MPa m1/2 to 0.97 MPa m1/2 measured by indentation fracture [Milekhine 2002]. However, Milekhine et al. used their measured E and H values (Tables 4.2 and 4.3) to calculate KIC, and their E and H values are lower than the E and H values found in this study and in the literature (Tables 4.2 and 4.3). Using the crack length, c, reported by Milekhine et al. and the E and H values from this study (117 GPa and 5.1 GPa, respectively), the recalculated values for Milekhine et al.’s KIC values would be 0.8 MPa m1/2, 0.9 MPa m1/2, and 1.1 MPa m1/2 using Eq. 2, which are in better agreement with the data from this study (Table 4.3). The KIC values of other thermoelectric materials include 1.12 MPa m1/2 to 1.35 MPa m1/2 for Bi2Te3 [Zhao 2008], 0.35 MPa m1/2 for PbTe-PbS 8% [Ni 2010], and 0.34 MPa m1/2 for LAST [Ren 2008]. The radial crack length was measured on six indentations of specimen Mg2Si-01. Slow crack growth was not observed when examining the average length of the two radial cracks per indentation (Fig. 4.4) as a function of time for six indentations for a period of days (Fig. 4.5). Two indentations had a crack pop-in within the first 5 min after indentation, which was not considered slow crack growth. The mean and standard deviation of the radial crack length measurement versus time are shown in Fig. 4.5 and Table 4.4, indicating that there was no statistically significant slow crack growth that 96 Crack Length (µm) Crack Length (µm) 110 100 90 80 (a) 70 1 10 100 1000 110 100 90 80 70 (b) 1 10 110 100 90 80 (c) 1 10 100 Crack Length (µm) Crack Length (µm) 110 1000 100 90 80 (d) 70 1 Time (min) 100 1000 110 Crack Length (µm) Crack Length (µm) 10 Time (min) 110 100 90 80 70 1000 Time (min) Time (min) 70 100 (e) 1 10 100 1000 100 90 80 (f) 70 1 10 100 1000 Time (min) Time (min) Figure 4.5. For specimen Mg2Si-01, plots of radial crack length versus time for six Vickers indentation crack systems loaded at 2.94 N. In figures (b) and (e), a radial crack popped in within 1 min to 5 min following the initial indentation event. The average crack length and one standard deviation variation is indicated by solid and dashed lines, respectively, for each indentation crack. The crack length may vary significantly between indentations, resulting in the uncertainties in the reported KIC values in Table 4.3. 97 Table 4.4. No significant slow crack growth occurred for the six Vickers indentation cracks listed below for specimen Mg2Si-01, where the radial crack lengths were monitored for up to 3 days (Fig. 4.5). For the radial crack length measurements performed over the entire time range, the measured crack lengths did not depart significantly from the mean length of each crack, as indicated by the coefficient of variation (CV, standard deviation/mean), where CV ranges from about 1.4% to 5%. Indent identification Mg2Si-01 (a) Mg2Si-01 (b) Mg2Si-01 (c) Mg2Si-01 (d) Mg2Si-01 (e) Mg2Si-01 (f) Mean crack length (µm) 106.5 78.2 107.3 97.5 86.5 86.2 Standard deviation (µm) 1.5 2.9 1.6 2.0 4.5 2.3 98 Coefficient of variation (CV) 0.014 0.037 0.014 0.020 0.051 0.027 occurred for Mg2Si for times up to 3 days (Fig. 4.5). There was no pattern of crack length growth with time outside of one standard deviation (Fig. 4.5). Differences in crack length between Figs. 4.5a–f are based on the individual indentations. With no pattern of crack length change with time, any variation in length measurements (Table 4.4) for an indentation may be attributed to the error of the optical measurement technique. 4.4 Conclusions The elastic moduli, hardness, fracture toughness, and slow crack growth were measured on powder- processed and PECS-sintered specimens of Mg2Si. The average E of 117.3 GPa, H of 5.3 GPa, and KIC of 1.3 MPa m1/2 obtained in this study agree well with the literature (Tables 4.2 and 4.3). No examination of the susceptibility of Mg2Si to slow crack growth has appeared in the literature prior to this study. In spite of a known reaction with water [Aesar], this reaction did not drive slow crack growth for the radial cracks observed in this study. The lack of significant slow crack growth is beneficial to the development and fabrication of Mg2Si, as microcracks induced by cutting or grinding during fabrication would not be expected to grow under normal atmospheric conditions. Mg2Si has several advantages as a potential thermoelectric material. It has low mass density and contains only nontoxic, inexpensive, and readily available elements. Thus, Mg2Si is a very viable candidate for widespread application as a thermoelectric material. ACKNOWLEDGEMENTS The authors acknowledge the financial support of the Department of Energy, “Revolutionary Materials for Solid State Energy Conversion Center,” an Energy 99 Frontiers Research Center funded by the US Department of Energy, Office of Basic Energy Sciences under Award Number DE-SC0001054. Jennifer Ni (under the Office of Naval Research Grant N00014-08-1-0613) assisted with the scanning electron microscopy of the specimen surfaces and assisted Jesse Giles with the hardness measurements. Financial support for Mr. Giles was provided through the Michigan Louis Stokes Alliance for Minority Participation (MI-LSAMP) Summer Undergraduate Research Academy (SURA), which was hosted in collaboration with the Michigan State University Summer Research Opportunity Program (MSU SROP). 100 REFERENCES 101 REFERENCES [Aesar] Aesar, A. Material Safety Data Sheet, Magnesium Silicide. http://www.alfa.com/content/msds/USA/45518.pdf. [Barsoum 2003] Barsoum, M.W. Fundamentals of Ceramics. Taylor & Francis Group, New York, 2003. [Case, Smyth, Hunter 1981] Case, E.D., Smyth, J.R., Hunter, O. Microcracking in Largegrain AI203. 51 (1981) 175–179. [Connally, Brown 1992] Connally, J.A., Brown, S.B. Slow crack growth in single-crystal silicon. Science 256 (1992) 1537–9. [Freiman, Wiederhorn, Mecholsky 2009] Freiman, S.W., Wiederhorn, S.M., Mecholsky, J.. J.J. Environmentally Enhanced Fracture of Glass: A Historical Perspective. Journal of the American Ceramic Society 92 (2009) 1371–1382. [Hazen 1976] Hazen, R. Effects of temperature and pressure on the cell dimension and Xray temperature factors of periclase. American Mineralogist 61 (1976) 266–271. [Lawn 1993] Lawn, B.R. Fracture of Brittle Solids. Cambridge University Press, New York, 1993. [Li, Gill, Varin 1993] Li, G., Gill, H., Varin, R. Magnesium silicide intermetallic alloys. Metallurgical transactions A 24 (1993) 2383–2391. [Mars, Ihou-Mouko, Pont, Tobola, Scherrer 2009] Mars, K., Ihou-Mouko, H., Pont, G., Tobola, J., Scherrer, H. Thermoelectric Properties and Electronic Structure of Biand Ag-Doped Mg2Si1−x Ge x Compounds. Journal of Electronic Materials 38 (2009) 1360–1364. [Michalsky, Freiman 1983] Michalsky, T.A., Freiman, S.W. A Molecular Mechanism for Stress Corrosion in Vitreous Silica. Journal of the American Ceramic Society 66 (1983) 284–288. [Milekhine, Onsøien, Solberg, Skaland 2002] Milekhine, V., Onsøien, M.I., Solberg, J.K., Skaland, T. Mechanical properties of FeSi (e), FeSi2 (za) and Mg2Si. Intermetallics 10 (2002) 743–750. [Ni, Case, Khabir, Stewart, Wu, Hogan, Timm, Girard, Kanatzidis 2010] Ni, J.E., Case, E.D., Khabir, K.N., Stewart, R.C., Wu, C.-I., Hogan, T.P., Timm, E.J., Girard, S.N., Kanatzidis, M.G. Room temperature Young’s modulus, shear modulus, Poisson's 102 ratio and hardness of PbTe–PbS thermoelectric materials. Materials Science and Engineering: B 170 (2010) 58–66. [Owen, Preston 1923] Owen, E.A., Preston, G.D. The atomic structure of two intermetallic compounds. Proceedings of the Physical Society of London 36 (1923) 341–348. [Ren, Case, Ni, Timm, Lara-Curzio, Trejo, Lin, Kanatzidis 2009] Ren, F., Case, E.D., Ni, J.E., Timm, E.J., Lara-Curzio, E., Trejo, R.M., Lin, C.-H., Kanatzidis, M.G. Temperature-dependent elastic moduli of lead telluride-based thermoelectric materials. Philosophical Magazine 89 (2009) 143–167. [Ren, Hall, Ni, Case, Sootsman, Mercouri 2008] Ren, F., Hall, B.D., Ni, J.E., Case, E.D., Sootsman, J., Mercouri, G. Mechanical Characterization of PbTe-based Thermoelectric Materials. 1044 (2008) . [Schmidt, Ni, Case, Sakamoto, Kleinow, Wing, Stewart, Timm 2010] Schmidt, R.D., Ni, J.E., Case, E.D., Sakamoto, J.S., Kleinow, D.C., Wing, B.L., Stewart, R.C., Timm, E.J. Room temperature Young’s modulus, shear modulus, and Poisson's ratio of Ce0.9Fe3.5Co0.5Sb12 and Co0.95Pd0.05Te0.05Sb3 skutterudite materials. Journal of Alloys and Compounds 504 (2010) 303–309. [Simmons, Wang 1971] Simmons, G., Wang, H. Single Crystal Elastic Constants and Calculated Aggregate Properties: A Handbook. The M.I.T. Press, Cambridge, Massachusetts, 1971. [Tani, Kido 2008] Tani, J., Kido, H. Lattice dynamics of Mg2Si and Mg2Ge compounds from first-principles calculations. Computational Materials Science 42 (2008) 531– 536. [Underwood, Colcord, Waugh 1968] Underwood, E.E., Colcord, A.R., Waugh, R.C. Ceramic Microstructures. Wiley, New York, 1968. [Wachtman, Cannon, Matthewson 2009] Wachtman, J.B., Cannon, W.R., Matthewson, M.J. Mechanical Properties of Ceramics. Wiley-Vch, Hoboken, N.J., 2009. [West, Hench 1998] West, J.K., Hench, L.L. The effect of environment on silica fracture: Vacuum, carbon monoxide, water and nitrogen. Philosophical Magazine A 77 (1998) 85–113. [Whitten, Chung, Danielson 1965] Whitten, W., Chung, P., Danielson, G. Elastic constants and lattice vibration frequencies of Mg 2 Si. Journal of Physics and Chemistry of Solids 26 (1965) 49–56. 103 [Wiederhorn 1968] Wiederhorn, S. Moisture assisted crack growth in ceramics. International Journal of Fracture Mechanics 4 (1968) 171–177. [Zhang, Han, Yan, Wang, Xia, Xu 2009] Zhang, C., Han, P., Yan, X., Wang, C., Xia, L., Xu, B. First-principles study of typical precipitates in creep resistant magnesium alloys. Journal of Physics D: Applied Physics 42 (2009) 125403. [Zhang, Rogl, Grytsiv, Puchegger, Koppensteiner, Spieckermann, Kabelka, Reinecker, Rogl, Schranz, Zehetbauer, Carpenter 2010] Zhang, L., Rogl, G., Grytsiv, A., Puchegger, S., Koppensteiner, J., Spieckermann, F., Kabelka, H., Reinecker, M., Rogl, P., Schranz, W., Zehetbauer, M., Carpenter, M.A. Mechanical properties of filled antimonide skutterudites. Materials Science and Engineering: B 170 (2010) 26–31. [Zhao, Zhang, Li, Zhou, Liu, Liu 2008] Zhao, L.-D., Zhang, B.-P., Li, J.-F., Zhou, M., Liu, W.-S., Liu, J. Thermoelectric and mechanical properties of nano-SiC-dispersed Bi2Te3 fabricated by mechanical alloying and spark plasma sintering. Journal of Alloys and Compounds 455 (2008) 259–264. 104 5 Mechanical properties of Mg2Si thermoelectric materials with the addition of 0 to 4 volume percent silicon carbide nanoparticles (SiCNP) Robert D. Schmidt, Xiaofeng Fan, Eldon D. Case, Chemical Engineering and Materials Science Department, Michigan State University, East Lansing, MI, 48824 To be submitted to: Journal of Materials Science Abstract For thermoelectric generators, the individual thermoelectric elements are subjected to a significant stresses under in-service conditions, due to thermal cycles, transients and gradients, as well as thermal expansion mismatch and externally applied mechanical stresses. Most thermoelectric (TE) materials are brittle and with low fracture toughness, typically no more than 1.5 MPa m1/2 and often less than 0.5 MPa m1/2. The combination of the stresses encountered in the device application environment and the materials’ low fracture toughness constitute a severe challenge to the viability of thermoelectric generators. The addition of silicon carbide nanoparticles (SiCNP) may provide a route to improving the fracture toughness for a wide range of thermoelectric materials. This study examines the mechanical properties, including elastic modulus, hardness and fracture toughness for 0 to 4 vol.% SiCNP incorporated into Mg2Si thermoelectric matrices. 5.1 Background Thermoelectric (TE) materials for solid state energy conversion have been extensively investigated in recent years, with the potential to recapture energy from waste 105 heat sources. Typically, the efficiency of a material is measured by use of the dimensionless figure of merit, ZT, (1) where S is the Seebeck coefficient, σ is the electrical conductivity, κ is the thermal conductivity, and T is temperature. Thermoelectric generators, particularly when used in waste heat applications, are subjected to multiple sources of stress, including stresses due to thermal gradients, thermal shock, thermal expansion mismatch between materials, and externally applied stresses. Waste heat applications typically involve a heat source, such as exhaust from an engine. Automotive engines are typically run with several accelerations or decelerations, and the waste heat produced will likewise involve several thermal cycles during operation. The waste heat source therefore is a source with thermal transients, thermal shock, and thermal gradients, which produce stresses in the TE material. These stresses require a detailed knowledge of the mechanical properties of the material to understand the material response to the stress, particularly the elastic properties and the fracture toughness. The elastic moduli are necessary to understanding the mechanical response to stresses, including stresses that cause fracture and the calculation of fracture toughness. In addition, stresses and flaws resulting from the manufacturing processes, such as scratches and surface microcracks during cutting or grinding, may result in fracture. A typical TE module has tens to hundreds of legs, all thermally in parallel, but electrically in series, creating a condition where each leg is subjected to the full thermal cycle, but the failure in any one leg of the TE material electrically breaks the circuit and renders an 106 entire module of a generator inoperable. Therefore, an understanding and control of the mechanical response and failure of the TE materials is critical to the reliable operation of a TE generator. The fracture toughness, KC, of thermoelectric materials is typically very low, less than 1.5 MPa m1/2, and often less than 0.5 MPa m1/2 [Schmidt 2012; Eilertsen 2013; Ren 2008; Ni 2010]. For comparison, window glass has a fracture toughness of about 0.75 MPa m1/2 [Wiederhorn 1969]. Adding particles such as silicon carbide nanoparticles (SiCNP) may improve the fracture toughness for a wide range of TE materials. Fracture toughness may be increased by the incorporation of second phases, additions such as particles or whiskers, microstructural changes or other mechanisms. For the purposes of this study, these may be grouped into intrinsic or extrinsic toughening mechanisms [Launey 2009]. Intrinsic mechanisms (such as crack blunting or crack deflection) change the fracture toughness by acting on the area ahead of the crack tip, by distributing the load, changing the interface properties, or other methods to “increase the microstructural resistance” of the material [Launey 2009]. Extrinsic mechanisms (such as fiber or grain bridging) work on the area behind the crack tip, redistributing the force to reduce the force acting at the crack tip [Launey 2009]. These differences have important implications in the potential for improving the fracture toughness overall, and specific implications for materials under fatigue conditions such as thermal cycling. Extrinsic mechanisms may, for a few cycles, significantly improve fracture toughness. For example, in alumina the addition of 20% SiC whiskers increased KC by a multiple of 3 [Becher 1991], and additions of 20% and 40% tetragonal zirconia increased 107 KC by a multiple of 3 and 5, respectively [Becher 1991]. Similar increases were observed in glass and mullite [Becher 1991]. These extrinsic toughening methods, however, are not desirable in a fatigue condition [Ritchie 1999; Launey 2009; Case 2012a; Case 2012b] and can be destructive to the material [Lathabai 1991; Bhowmick 2007; Bhowmick 2009]. Bridging, for example, has been observed to be defeated by fatigue and can even be destructive in alumina [Knehans 1982; Lathabai 1991] and silicon [Bhowmick 2009]. When a bridge is defeated, the debris may fall into the crack and act as a wedge and extend the crack. Intrinsic toughening mechanisms, including crack deflection, crack blunting, crack branching or bowing, are not defeated by thermal fatigue [Launey 2009]. Therefore, intrinsic toughening mechanisms acting ahead of the crack tip are the desired type of toughening in fatigue [Launey 2009; Case 2012a; Case 2012b] as would be the case for a TE device in waste heat recovery that is subjected to multiple thermal cycles. The addition of nanoparticles can lead to intrinsic toughening. However, compared to large increases in KC that are possible with extrinsic toughening, intrinsic toughening offers smaller but significant increases in KC (Table 5.1). In alumina, additions of 5-20% SiC nanoparticles increases the relative KC by a factor of 1.16 [PerezRigueiro 1998; Parchovianský 2013]. In Bi2Te3, a TE material, additions of 0.1 to 0.5 vol% SiC nanoparticles increased relative KC by a factor of 1.18 [Zhao 2008]. Mg2Si may be used to examine several toughening mechanisms. Literature values of ZT show Mg2Si to be a good TE material (Table 5.2), with ZT of 0.86 at 862 K for (Mg2Si:Bi = 1:0.02) [Tani 2005] and 0.97 at 873 K for 0.5 at.% Sb doped Mg2Si + 5 wt. % Ni [Hayatsu 2012]. With values of ZT from 0.23 to 0.97 when doped with 2% or less 108 Table 5.1. The fracture toughness, KC, has been improved in brittle materials, including the thermoelectric Bi2Te3, by the addition of SiC nanoparticles. Matrix KC SiC addition KC Relative Reference 1/2 1/2 material (MPa·m ) (MPa·m ) change Bi2Te3 1.14 0.1 vol % 1.35 1.18 [Zhao 2008] 0.5 vol % 1.33 1.17 1.0 vol % 1.19 1.04 a a Al2O3 3.7 to 3.8 5 vol. % 3.6 0.96 a [Pérez-Rigueiro 3.0 to 3.3 b 20 vol. % a 3.4 to 3.6 0.93 a 1988] b b 5 vol. % 3.6 1.14 20 vol. % b 3.5 to 3.7 1.14 b Al2O3 5.0 ± 0.3 5 vol. % 5.4 to 5.6 1.10 [Parchovianský 10 vol. % 5.6 to 6.0 1.16 2013] 15 vol. % 5.2 to 5.3 1.05 20 vol. % 5.1 to 5.3 1.04 a Specimen tested by single-edged notched beam b Specimen tested by Vickers indentation Table 5.2. Doped Mg2Si-based thermoelectric materials with ZT near 1 have been reported. These reasonable ZT values for a thermoelectric material support the use of Mg2Si as a model system. Composition Maximum ZT Temperature for Reference maximum ZT (K) Mg2Si:Bi = 1:0.02 0.86 862 [Tani 2005] Mg2Si:Sb = 1:0.005 0.97 873 [Hayatsu 2012] + 5 wt.% Ni Mg2Si0.9Sn0.1:Al = 0.68 864 [Tani 2008] 1:0.02 Mg2Si:Y = 1:0.006 0.23 600 [Meng 2011] (2000 ppm of Y) Mg2Si:Mg2Pb = 1:0.02 0.56 873 [Muthiah 2013] Mg2Si:Sb = 1:0.02 0.56 862 [Tani 2007] 109 of dopant [Tani 2005; Hayatsu 2012; Tani 2008; Meng 2011; Muthiah 2013; Tani 2007], Mg2Si may be used as a suitable TE system to test additions of SiCNP. As reported on the MSDS, Mg2Si reacts with water [Aesar], which requires handling without water or water-based solutions. However, this reaction is not a problem for crack growth and mechanical integrity when the material is handled in dry air [Schmidt 2012]. Mg2Si is a suitable material for examination of fracture properties by Vickers indentation [Schmidt 2012]. To understand the mechanical response to the stresses, as well as the potential for fracture, the mechanical properties of the TE material must be not only understood, but optimized to properly withstand the thermal fatigue environment typical of many thermoelectric generator applications. 5.2 Experimental Procedure In this study, silicon carbide nanoparticles, SiCNP, were added to Mg2Si powders which were then processed to produce Mg2Si- SiCNP composites. 5.2.1 Materials and Specimen Preparation Powder was produced from lump Mg2Si (45518, 3-12 mm pieces, Alfa Aesar, Ward Hill MA), ground and sieved through a 53 μm sieve, then milled by either (i) a planetary ball mill (PM) using an alumina-lined mill jar with 10 mm diameter alumina grinding media at 150 RPM for 3 h, or (ii) a vibratory mill (VM) with a WC lined jar, two 11.2 mm spheres and four 7.9 mm spheres, and a Viton gasket, sealed in the glove box with an additional layer of electrical tape. The PM processed Mg2Si powders were milled in an argon-filled glove box and the VM processed powders were milled in a sealed, argon-filled jar. 110 Specimens were densified by pulsed electric current sintering (SPS Model 10-3, Thermal Technology LLC, Santa Rosa, CA) in a 12.7 mm diameter graphite die. Sintering was performed at 800°C and 50 MPa for 5 min, and with ramp rates of 100°C/min and 45 MPa/min. 5.2.2 Elasticity Measurements Elastic moduli were measured by resonant ultrasound spectroscopy (RUS). In RUS, the specimen is placed on a tripod of transducers. One transducer was swept through a range of frequencies and the mechanical resonance frequencies are picked up by the other two transducers. The resonances were fit to a model for the geometry, mass and dimensions of the specimen to determine the elastic moduli. Additional details of the RUS procedure are provided elsewhere [Ni 2010; Schmidt 2010; Migliori 1997]. 5.2.3 Hardness and toughness measurements Prior to indentation, each specimen was mounted with thermoplastic (CrystalBond 509, EMS, Hatfield, PA) onto an aluminum stub, then polished with a series of diamond compounds with grit size from 35 µm to 1 µm. Cleaning of the specimen between polishing grit sizes was performed by hand with ethanol rather than water to lessen any surface reaction between the Mg2Si and water [Aesar]. Hardness and fracture toughness for each of the specimens were measured by Vickers indentation. Vickers hardness, H, is calculated by the equation, H= 1.8544 F ( 2a ) 2 (2) 111 where F is the indentation load, 2a is the diagonal impression length, and ζ is a correction factor, set at 0.95 to 0.97 based on indentations of a steel standard calibration block (Yamamoto Scientific Tools Lab Co. LTD, Chiba, Japan). The fracture toughness, KC, of a material may be estimated by measuring the radial crack length, c, of a Vickers indentation, by the equation, KC   E H 1 2 P (3) c3 2 where ξ is a dimensionless constant, set as 0.016, E is the Young’s modulus, H is the hardness, and P is the applied load [Wachtman 2009]. 5.2.4 Microscopy The specimens were examined by scanning electron microscope (SEM) at a working distance of either 8 mm or 15 mm, and at an accelerating voltage of 15 kV (JEOL 6610LV or JSM-7500F, JEOL Ltd., Japan). The average grain size (GS) of each of the sintered specimens was determined using the linear intercept procedure (ASTM E112) with a minimum of 200 intercepts per image, and a stereographic projection factor of 1.5 [ASTM-Standard-E112-13 2014]. Surface details were examined with secondary electron imaging (SEI) and elemental contrasts were examined by backscatter electron imaging (BEI). 5.3 Results and Discussion 5.3.1 Microstructural analysis Prior to sintering, both the PM and the VM powders were examined for size range and morphology. For the PM powder, there was a wide distribution from sub-micron 112 Figure 5.1. Planetary milled Mg2Si powder exhibited typical particle sizes of sub-micron to 5 μm in SEM. The powder has a surface area of 4.4 m2/g measured by BET, or approximately 0.7 μm average particle size. Figure 5.2. Vibratory milled Mg2Si powder exhibited typical particle sizes of 0.2 μm to 2 μm in SEM. 113 particles to 5 μm particles or greater, average particle size by BET was 0.7 μm (Figure 5.1). To incorporate SiCNP with a vendor specified average particle size of 45-55 nm (44646, Lot number E15T018, Alfa Aesar) into the PM powder, the PM milled Mg2Si and SiCNP were planetary ball milled at 120 RPM for 3 h in an argon-filled glove box with the same mill jar and media used to mill the Mg2Si powder. For the VM powder, there was a relatively uniform distribution of powder, 0.2 μm to 2 μm in size (Figure 2). The SiCNP (44646, Lot number E15T018, Alfa Aesar) were incorporated into the VM powder simultaneously with milling the sieved powder. The original PM powder average particle size of 0.7 µm (Figure 1)was smaller than the average GS of the sintered PM specimens of 2.0 µm to 3.9 µm (Table 5.3), indicating some grain growth during sintering (Figure 3a – c). No BET particle size measurement was performed on the VM powder to compare to GS, however the average GS of the VM specimens of 0.4 μm to 0.8 μm (Table 5.3) is consistent with only limited grain growth from the original VM powder (Figure 5.2) to the sintered specimens (Figure 5.3d). 5.3.1.1 Starting material density and microstructure All sintered specimens in this study except one have a density of 2.00 to 2.06 g/cm3. The one specimen with a density of 1.93 g/cm3, VM-0SiC-02, is mentioned individually, but any averages or trends in mechanical properties excluded VM-0SiC-02 as an outlier. Most of the sintered specimen densities are higher than theoretical density for Mg2Si of 2.00 g/cm3 reported in the literature [Whitten 1965]. The higher density is 114 Table 5.3. Specimens in this study were either milled by planetary ball mill (PM) or vibratory mill (VM), with up to 4 vol% SiCNP additions, then sintered by pulsed electric current sintering to produce specimens with a density (ρ) of 2.00 g/cm3 or greater for all specimens except VM-0SiC-2. The average grain size (GS) by the lineal intercept method are a function of the milling method, and not a function of SiCNP additions. SiCNP ρ GS Specimen vol% addition (g/cm3) (µm) PM-0SiC 0.0 2.03 2.4 PM-0.5SiC 0.5 2.02 2.0 PM-1SiC 1.0 2.02 2.0 PM-1.5SiC 1.5 2.01 2.1 PM-2SiC 2.0 2.06 1.3 PM-3SiC 3.0 2.04 3.9 PM-4SiC 4.0 2.02 3.6 VM-0SiC-1 0.0 2.00 0.6 VM-0SiC-2 0.0 1.93 0.4 VM-0.5SiC 0.5 2.01 0.6 VM-1SiC 1.0 2.03 0.7 VM-1.5SiC 1.5 2.03 0.8 VM-2SiC 2.0 2.03 0.6 115 Figure 5.3. For Mg2Si fracture surfaces, transgranular fracture dominate in all specimens. The PM specimens (a-c) with varying amounts of SiCNP addition did not show any appreciable difference in grain size, although changing to VM processing significantly reduced the grain size (d). Note the difference in scale between the PM images (a-c) and the VM image (d). 116 because of the presence of MgO within the material. The starting material as received from the vendor has been shown to contain approximately 8 vol% of MgO, and likely to have a composite theoretical density near 2.12 g/cm3 [Schmidt 2012] (see section 5.3.2). Previous examination of the SiCNP material has shown that the as-received nanopowder from two vendors (Alfa Aesar and Nanostructured & Amorphous Materials Inc.) consists primarily of particles, approximately 50 nm in diameter, agglomerated into clusters of 100 nm to 20 µm in diameter [Schmidt 2013b]. These agglomerates do not appreciably break apart with milling [Schmidt 2013b]. In this study, the SiCNP material was from one of the two previously examined vendors (Alfa Aesar). 5.3.2 Elasticity results The elasticity for specimens with SiCNP, excluding the outlier specimen VM0SiC-02 with density of 1.93 g/cm3, is not a function of SiCNP addition up to 4 vol%, nor powder processing methodology (Figure 5.4, Table 5.4). Average Young’s modulus, E, for the PM specimens with 0 to 4 vol% SiCNP is 113.3 ± 4.0 GPa, and 111.1 ± 2.3 GPa for VM specimens (Figure 5.4a – b). The small variations in elastic moduli observed are not observed to be a function of the SiCNP addition (Figure 5.4), rather a function of the limited differences in porosity among specimens (Figure 5.5). 5.3.2.1 Elasticity and SiCNP addition If the SiCNP addition does not react with the Mg2Si matrix, the elastic modulus of the composite material with addition of up to 4 vol% SiCNP may be modeled as a mixture of two independent materials. In a study of SnTe with 0 vol% to 4 vol% SiCNP addition [Schmidt 2013b], the Hashin model [Hashin 1962; Bedolla 2012; Couturier 1997] 117 110 3 Density (g/cm ) 2.070 100 2.050 2.030 2.010 90 2.000 0 2 Young's modulus (GPa) Young's modulus (GPa) 120 (a) Planetary milled 120 (b) Vibratory milled 110 3 Density (g/cm ) 2.070 100 2.050 2.030 2.010 90 2.000 4 0 Vol. % SiCNP 55 (d) Vibratory milled 3 Density (g/cm ) 2.070 40 2.040 2.020 35 2.010 2.000 30 2 Shear modulus (GPa) Shear modulus (GPa) 50 0 4 Vol. % SiCNP 55 (c) Planetary milled 45 2 4 50 45 3 Density (g/cm ) 2.070 40 2.050 2.030 35 2.010 2.000 30 0 2 4 Vol. % SiCNP Vol. % SiCNP Figure 5.4. The Young’s modulus (a, b) and shear modulus (c, d) of Mg2Si varied primarily due to small variations in density. The variation in vol. % SiCNP did not significantly affect the moduli, regardless of if the specimens were planetary milled (a, c) or vibratory milled (b, d). Young's Modulus (GPa) 140 PM specimens VM specimens 130 120 110 100 90 0.00 0.02 0.04 0.06 0.08 0.10 Porosity (Vol. Fraction) Figure 5.5. The Young’s modulus, E, decreases linearly with porosity for the set of 13 specimens in this study. 118 Table 5.4. The Young’s modulus, E, shear modulus, G, and density, ρ, for the specimens in this study, as compared to the range of E, G and ρ in a previous study on three Mg2Si specimens produced by the same vendor. Specimen E (MPa) G (MPa) ρ (g/cm3) Reference PM-0SiC 113.6 ± 0.3 48.3 ± 0.1 2.03 This study PM-0.5SiC 112.9 ± 0.3 47.5 ± 0.1 2.02 This study PM-1SiC 111.2 ± 0.3 46.1 ± 0.1 2.02 This study PM-1.5SiC 105.8 ± 0.4 43.3 ± 0.1 2.01 This study PM-2SiC 118.1 ± 0.2 50.3 ± 0.1 2.06 This study PM-3SiC 116.7 ± 0.2 49.7 ± 0.1 2.04 This study PM-4SiC 114.5 ± 0.6 49.7 ± 0.2 2.02 This study VM-0SiC-1 110.4 ± 0.2 46.9 ± 0.1 2.00 This study VM-0SiC-2 90.4 ± 0.3 37.3 ± 0.1 1.93 This study VM-0.5SiC 107.6 ± 0.4 44.7 ± 0.1 2.01 This study VM-1SiC 113.4 ± 0.2 48.3 ± 0.1 2.03 This study VM-1.5SiC 113.0 ± 0.2 48.1 ± 0.1 2.03 This study VM-2SiC 111.3 ± 0.3 47.5 ± 0.1 2.03 This study Mg2Si 116.9 – 48.92 – 50.09 2.07 – [Schmidt 2012] 117.7 2.09 119 described relatively well the resulting change in Young’s modulus where the Hashin model can be written as  E V  Er Vr  1  EC  Em  m m  ErVm  Em Vr  1 (4) where EC is the Young’s modulus of the composite, Em is the Young’s modulus of the matrix material, Er is the Young’s modulus of the reinforcing phase, and Vm and Vr are the volume fraction of the matrix and the reinforcing phase, respectively. Applying the Hashin model (equation (4)) to the materials in this study, using the Young’s modulus of the reinforcing phase SiCNP, Er, at 450 GPa [Schreiber 1966; Carnahan 1968] and the Em of 112 GPa the Mg2Si measured in this study, the EC for the composite with the maximum 4 vol% SiCNP is 117.5 GPa, which is similar to the measured E of 114.5 ± 0.6 GPa from specimen PM-4SiC (Figure 5.4a). 5.3.2.2 Elasticity and porosity Small variations in density of the specimens in this study, from 2.00 to 2.06 g/cm3, is likely the dominant source of variation in E measured in this study (Figure 5.4a – b, Figure 5.5). Typically, for small variations in porosity, P, the change in E with porosity may be modeled by the empirical equation, E  E0 1  bE P  (5) where E0 is the Young’s modulus of a fully dense material and bE is a materialdependent constant, typically between 2 and 6 for a wide range of materials [Rice 1998]. For TE materials YbAl3 [Schmidt 2013a] and lead-antimony-silver-tellurium (LAST) [Ni 2009], the experimentally determined values of bE were 2.34 and 3.5, respectively . Assuming the matrix material contains 8% MgO [Schmidt 2012] (density 3.58 g/cm3 120 [Anderson 1966; Bogardus 1965]) and 92% Mg2Si (density 2.00 g/cm3 [Whitten 1965]), the composite theoretical densities (matrix density, 2.126 g/cm3, SiCNP density, 3.16 g/cm3 [Aesar]) were calculated. The volume fraction porosity, P, was calculated for the composite materials. The E versus P for the full set of 13 composite specimens in this study were fit to equation (5), with the bE was 3.5 ± 0.5 and the E0 was 138 ± 5 GPa (Figure 5.5). The linear relationship indicates that the E for the 14 specimens is a function of P. The error on the E0 and bE are relatively large because of the restricted range of P for the specimens this study. A previous study on Mg2Si from lump material purchased from the same vendor and processed by PM measured E of 116.9 GPa to 117.7 GPa [Schmidt 2012] (Table 5.4). The small (~4%) difference in elastic moduli between this study and the previous study is likely due to the slightly higher density specimens in the previous study, 2.07 g/cm3 to 2.09 g/cm3 (Table 5.4). 5.3.3 Hardness and Toughness Results Hardness, H, by Vickers indentation for specimens with SiCNP, excluding one outlier, was relatively insensitive to vol% SiCNP, from 0 % to 4 vol% SiCNP, indentation load, from 1.96 N to 4.9 N, or powder processing technique, averaging 4.76 ± 0.37 GPa for the PM specimens (Figure 5.6a) and 4.83 ± 0.15 GPa for the VM specimens (Figure 5.6b). The outlier is specimen VM-0SiC-02 with a density of 1.93 g/cm3, lower than the 2.00 g/cm3 to 2.06 g/cm3 density of the other 12 specimens in this study. The H of the outlier VM-0SiC-02 specimen is indicated by the open symbol in Figure 5.6b. The fracture toughness, KC, is dependent on the powder processing conditions. The KC of the PM composite reached a maximum at 1 vol% SiCNP, an increase of about 121 (a) 1.96 N load 2.94 N load 4.9 N load 7 Hardness (GPa) 6 5 4 3 2 1/2 1.5 (c) Planetary milled 1.96 N load 2.94 N load 4.9 N load 1.2 0.9 0.6 0.3 0.0 0 2 Vol. % SiCNP 4 Vibratory milled 1.96 N load 2.94 N load 4.9 N load 5 4 3 4 Vol. % SiCNP (b) 6 0 2 4 Vol. % SiCNP 1/2 0 Fracture Toughness (MPa m ) Planetary milled Fracture Toughness (MPa m ) Hardness (GPa) 7 1.5 (d) Vibratory milled 1.96 N load 2.94 N load 4.9 N load 1.2 0.9 0.6 0.3 0.0 0 2 4 Vol. % SiCNP Figure 5.6. The hardness, (a) and (b) is not a function of the milling procedure or the vol% SiCNP, but less scatter was observed in the (b) vibratory milled specimens than the (a) planetary ball milled specimens. The fracture toughness exhibited a maximum at 1 vol% SiCNP for the (c) planetary ball milled specimens, but the fracture toughness is not a function of vol% SiCNP for the (d) vibratory milled specimens. Open symbols in (b) and (d) indicate a specimen with lower density of 1.93 g/cm3, relative to the 2.00 to 2.06 g/cm3 for all other specimens in this study. 122 33% over the unreinforced PM material, and is relatively insensitive to SiCNP addition between 1 vol% and 2 vol% SiCNP addition (Figure 5.6c). In contrast, the VM composite material was relatively insensitive to any additions of up to 2 vol% SiCNP (Figure 5.6d). The VM material also exhibited a much smaller variation in measured KC, with the coefficient of variation, CV, averaging 0.08 and ranging from 0.04 to 0.14 (Figure 5.6d). The CV of the PM material averages 0.16 and ranging from 0.11 to 0.34. The smaller CV observed in the VM specimens relative to the PM specimens may be related to the smaller GS of the VM specimens, leading to a more uniform crack length and possibly more uniform distribution of bridging and stalling. The smaller CV is particularly important when considering the distribution of strength, and the related probability of failure at a given load. The strength of the material that may be used safely in a design depends upon the scatter in strength [Wachtman 2009], and the smaller CV for the strength values in the VM specimens indicates a smaller scatter of strength for VM specimens than PM specimens. In this study, KC is not a function of GS for the Mg2Si with no SiCNP addition (Figure 5.6c – d). This result is expected for cubic materials such as Mg2Si, because KC is largely independent of GS [Wachtman 2009], and specifically for Vickers indentation crack length, “there should be a rather weak dependence of crack length on GS” [Wachtman 2009, pg 217]. However, the effect of SiCNP addition on KC was influenced by the powder processing technique. The difference in size between the added SiCNP particles and the Mg2Si matrix may play a role in the behavior of KC. In a study on fracture toughness on TE material specimens sintered from a blend of nanoparticles with microparticles of the 123 Figure 5.7. Crack bridging, in PM Mg2Si was commonly observed in radial cracks for all the Mg2Si specimens in this study. 124 Figure 5.8. Crack bridges were commonly observed in VM specimens regardless of SiCNP addition. Crack bridging in radial cracks was not eliminated by reducing grain size through VM processing. Spotting is from oil residue on the surface of the specimen. 125 same composition, Co4Sb11.5Te0.5, the KC increased with additions of nanoparticles [Duan 2014]. Similar to the KC of Co4Sb11.5Te0.5 [Duan 2014], this study indicates that the toughening due to addition of SiCNP may depend on GS of the matrix Mg2Si relative to the SiCNP. 5.3.3.1 Crack bridging and toughness Crack bridging and stalling was observed on all of the different Mg2Si specimens, both with and without additions and regardless of processing by PM (Figure 5.7) or VM (Figure 5.8). Both bridging and stalling are extrinsic toughening mechanisms, and not beneficial to KC in a fatigue loading condition such as thermal cycling [Lathabai 1991; Ritchie 1999]. In some cases, these may be controlled for through processing. Bridging in monophase alumina material was eliminated by reducing the GS [Lathabai 1991], although changing the powder processing of the Mg2Si in this study did not produce a similar result. 5.4 Conclusions Fracture toughness, KC, of Mg2Si increases for PM material by 33% with the addition of 1 – 2 vol% SiCNP, but KC is independent of SiCNP addition for VM material. As is typical of cubic materials such as Mg2Si, KC is independent of GS for specimens with no addition of SiCNP. The coefficient of variation for the KC of the VM samples was smaller, 0.08, than for the PM samples, 0.16, indicating a likely smaller scatter of specimen strength for VM specimens than the PM specimens. ACKNOWLEDGEMENTS The authors acknowledge the financial support of the Department of Energy, Revolutionary Materials for Solid State Energy Conversion Center, an Energy Frontiers 126 Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under award number DE-SC0001054. 127 REFERENCES 128 REFERENCES [Aesar] Aesar, A. Material Safety Data Sheet, Magnesium Silicide. http://www.alfa.com/content/msds/USA/45518.pdf. [Anderson, Andreatch 1966] Anderson, O.L., Andreatch, P. Pressure Derivatives of Elastic Constants of Single-Crystal MgO at 23° and -195.8°C. Journal of the American Ceramic Society 49 (1966) 404–409. [ASTM-Standard-E112-13 2014] ASTM-Standard-E112-13. Standard Test Methods for Determining Average Grain Size. In: ASTM Volume 03.01 Metals Mechanical Testing; Elevated and Low Temperature Tests; Metallography. ASTM International, West Conshohocken, PA, 2014. [Becher 1991] Becher, P.F. Microstructural Design of Toughened Ceramics. Journal of the American Ceramic Society 74 (1991) 255–269. [Bedolla, Lemus-Ruiz, Contreras 2012] Bedolla, E., Lemus-Ruiz, J., Contreras, A. Synthesis and characterization of Mg-AZ91/AlN composites. Materials and Design 38 (2012) 91–98. [Bhowmick, Cha, Jung, Lawn 2009] Bhowmick, S., Cha, H., Jung, Y.-G., Lawn, B.R. Fatigue and debris generation at indentation-induced cracks in silicon. Acta Materialia 57 (2009) 582–589. [Bhowmick, Meléndez-Mart nez, awn 2007] Bhowmick, S., Meléndez-Mart nez, . ., Lawn, B.R. Bulk silicon is susceptible to fatigue. Applied Physics Letters 91 (2007) 201902. [Bogardus 1965] Bogardus, E.H. Third-Order Elastic Constants of Ge, MgO, and Fused SiO2. Journal of Applied Physics 36 (1965) 2504. [Carnahan 1968] Carnahan, R. Elastic Properties of Silicon Carbide. Journal of the American Ceramic Society 51 (1968) 223–224. [Case 2012a] Case, E.D. Thermo-mechanical properties of thermoelectric materials. In: D.M. Rowe, ed., Thermoelectrics and its Energy Harvesting: Modules, Systems and Applications. CRC Press, Boca Raton, Florida, 2012, 581. [Case 2012b] Case, E.D. Thermal Fatigue and Waste Heat Recovery via Thermoelectrics. Journal of Electronic Materials 41 (2012) 1811–1819. 129 [Couturier, Ducret, Merle, Disson, Joubert 1997] Couturier, R., Ducret, D., Merle, P., Disson, J.P., Joubert, P. Elaboration and characterization of a metal matrix composite: Al/AlN. Journal of the European Ceramic Society 17 (1997) 1861–1866. [Duan, Zhai, Ding, Xu, Li, Liu, Li, Zhang 2014] Duan, B., Zhai, P., Ding, S., Xu, C., Li, G., Liu, L., Li, P., Zhang, Q. Effects of Nanoparticle Size on the Thermoelectric and Mechanical Properties of Skutterudite Nanocomposites. Journal of Electronic Materials 43 (2014) 2115–2120. [Eilertsen, Subramanian, Kruzic 2013] Eilertsen, J., Subramanian, M. a., Kruzic, J.J. Fracture toughness of Co4Sb12 and In0.1Co4Sb12 thermoelectric skutterudites evaluated by three methods. Journal of Alloys and Compounds 552 (2013) 492–498. [Hashin 1962] Hashin, Z. The Elastic Moduli of Heterogeneous Materials. Journal of Applied Mechanics 29 (1962) 143–150. [Hayatsu, Iida, Sakamoto, Kurosaki, Nishio, Kogo, Takanashi 2012] Hayatsu, Y., Iida, T., Sakamoto, T., Kurosaki, S., Nishio, K., Kogo, Y., Takanashi, Y. Fabrication of large sintered pellets of Sb-doped n-type Mg2Si using a plasma activated sintering method. Journal of Solid State Chemistry 193 (2012) 161–165. [Knehans, Steinbrech 1982] Knehans, R., Steinbrech, R. Memory effect of crack resistance during slow crack growth in notched AI203 bend specimens. (1982) 327– 329. [Lathabai, Rodel, Lawn 1991] Lathabai, S., Rodel, J., Lawn, B.R. Cyclic Fatigue from Frictional Degradation at Bridging Grains in Alumina. Journal of the American Ceramic Society 74 (1991) 1340–1348. [Launey, Ritchie 2009] Launey, M.E., Ritchie, R.O. On the Fracture Toughness of Advanced Materials. Advanced Materials 21 (2009) 2103–2110. [Meng, Fan, Chen, Munir 2011] Meng, Q.S., Fan, W.H., Chen, R.X., Munir, Z. a. Thermoelectric properties of Sc- and Y-doped Mg2Si prepared by field-activated and pressure-assisted reactive sintering. Journal of Alloys and Compounds 509 (2011) 7922–7926. [Migliori, Sarrao 1997] Migliori, A., Sarrao, J.L. Resonant Ultrasound Spectroscopy: Applications to Physics, Materials Measurements, and Nondestructive Evaluation. Wiley-Vch, New York, 1997. [Muthiah, Pulikkotil, Srivastava, Kumar, Pathak, Dhar, Budhani 2013] Muthiah, S., Pulikkotil, J., Srivastava, a. K., Kumar, A., Pathak, B.D., Dhar, A., Budhani, R.C. Conducting grain boundaries enhancing thermoelectric performance in doped Mg2Si. Applied Physics Letters 103 (2013) 053901. 130 [Ni, Ren, Case, Timm 2009] Ni, J., Ren, F., Case, E., Timm, E. Porosity dependence of elastic moduli in LAST (lead–antimony–silver–tellurium) thermoelectric materials. Materials Chemistry and Physics(2009) . [Ni, Case, Khabir, Stewart, Wu, Hogan, Timm, Girard, Kanatzidis 2010] Ni, J.E., Case, E.D., Khabir, K.N., Stewart, R.C., Wu, C.-I., Hogan, T.P., Timm, E.J., Girard, S.N., Kanatzidis, M.G. Room temperature Young’s modulus, shear modulus, Poisson's ratio and hardness of PbTe–PbS thermoelectric materials. Materials Science and Engineering: B 170 (2010) 58–66. [Parchovianský, Galusek, Sedláček, Švančárek, Kašiarová, Dusza, Šajgal k 2013] Parchovianský, M., Galusek, D., Sedláček, ., Švančárek, P., Kašiarová, M., Dusza, ., Šajgal k, P. Microstructure and mechanical properties of hot pressed Al2O3/SiC nanocomposites. Journal of the European Ceramic Society 33 (2013) 2291–2298. [Perez-Rigueiro, Pastor, Llorca 1998] Perez-Rigueiro, J., Pastor, J., Llorca, J. Revisiting the mechanical behavior of alumina/silicon carbide nanocomposites. Acta materialia 46 (1998) . [Ren, Case, Timm, Schock 2008] Ren, F., Case, E.D., Timm, E.J., Schock, H.J. Hardness as a function of composition for n-type LAST thermoelectric material. Journal of Alloys and Compounds 455 (2008) 340–345. [Rice 1998] Rice, R.W. Porosity of Ceramics. Marcel Dekker, New York, 1998. [Ritchie 1999] Ritchie, R. Mechanisms of fatigue-crack propagation in ductile and brittle solids. International Journal of Fracture(1999) 55–83. [Schmidt, Case, Giles, Ni, Hogan 2012] Schmidt, R.D., Case, E.D., Giles, J., Ni, J.E., Hogan, T.P. Room-Temperature Mechanical Properties and Slow Crack Growth Behavior of Mg2Si Thermoelectric Materials. Journal of Electronic Materials 41 (2012) 1210–1216. [Schmidt, Case, Lehr, Morelli 2013a] Schmidt, R.D., Case, E.D., Lehr, G.J., Morelli, D.T. Room temperature mechanical properties of polycrystalline YbAl3, a promising low temperature thermoelectric material. Intermetallics 35 (2013) 15–24. [Schmidt, Case, Ni, Trejo, Lara-Curzio, Korkosz, Kanatzidis 2013b] Schmidt, R.D., Case, E.D., Ni, J.E., Trejo, R.M., Lara-Curzio, E., Korkosz, R.J., Kanatzidis, M.G. High-temperature elastic moduli of thermoelectric SnTe1±x – y SiC nanoparticulate composites. Journal of Materials Science 48 (2013) 8244–8258. [Schmidt, Ni, Case, Sakamoto, Kleinow, Wing, Stewart, Timm 2010] Schmidt, R.D., Ni, J.E., Case, E.D., Sakamoto, J.S., Kleinow, D.C., Wing, B.L., Stewart, R.C., Timm, E. . Room temperature Young’s modulus, shear modulus, and Poisson's ratio of 131 Ce0.9Fe3.5Co0.5Sb12 and Co0.95Pd0.05Te0.05Sb3 skutterudite materials. Journal of Alloys and Compounds 504 (2010) 303–309. [Schreiber, Soga 1966] Schreiber, E., Soga, N. Elastic Constants of Silicon Carbide. Journal of the American Ceramic Society 49 (1966) 342. [Tani, Kido 2005] Tani, J., Kido, H. Thermoelectric properties of Bi-doped Mg2Si semiconductors. Physica B: Condensed Matter 364 (2005) 218–224. [Tani, Kido 2007] Tani, J., Kido, H. Thermoelectric properties of Sb-doped Mg2Si semiconductors. Intermetallics 15 (2007) 1202–1207. [Tani, Kido 2008] Tani, J., Kido, H. Thermoelectric properties of Al-doped Mg2Si1−xSnx (x≦0.1). Journal of Alloys and Compounds 466 (2008) 335–340. [Wachtman, Cannon, Matthewson 2009] Wachtman, J.B., Cannon, W.R., Matthewson, M.J. Mechanical Properties of Ceramics. Wiley-Vch, Hoboken, N.J., 2009. [Whitten, Chung, Danielson 1965] Whitten, W., Chung, P., Danielson, G. Elastic constants and lattice vibration frequencies of Mg 2 Si. Journal of Physics and Chemistry of Solids 26 (1965) 49–56. [Wiederhorn 1969] Wiederhorn, S. Fracture surface energy of glass. Journal of the American Ceramic Society 52 (1969) 99–105. [Zhao, Zhang, Li, Zhou, Liu, Liu 2008] Zhao, L.-D., Zhang, B.-P., Li, J.-F., Zhou, M., Liu, W.-S., Liu, J. Thermoelectric and mechanical properties of nano-SiC-dispersed Bi2Te3 fabricated by mechanical alloying and spark plasma sintering. Journal of Alloys and Compounds 455 (2008) 259–264. 132 6 Influence of silver nanoparticle addition, porosity and processing technique on the mechanical properties of Ba0.3Co4Sb12 skutterudites Robert D. Schmidt1, Eldon D. Case1, Zayra Lobo1, Travis R. Thompson1, Jeffrey S. Sakamoto1, Xiao-Yuan Zhou2, Ctirad Uher2 1 Chemical Engineering and Materials Science Department, Michigan State University, East Lansing, MI, 48824 2 Physics Department, University of Michigan, Ann Arbor, MI, 48109 Published in: Journal of Materials Science (2014) DOI:10.1007/s10853-014-8427-5. Abstract The thermoelectric skutterudite Ba0.3Co4Sb12 is a promising candidate for waste heat recovery applications. Recently, it was demonstrated that the addition of silver nanoparticles (AgNP) to Ba0.3Co4Sb12 increases both the thermoelectric figure of merit and electrical conductivity. This study is the first to examine the effect of AgNP addition on the material’s mechanical properties. This study also found that the Young’s modulus, E, shear modulus, G, and bulk modulus, B decreased linearly with increasing volume fraction porosity, P. Resonant ultrasound spectroscopy (RUS) was employed to measure the elastic moduli and Vickers indentation was used to determine the hardness, H, and fracture toughness, KC. Trends in the mechanical properties as a function of grain size, porosity and the AgNP are discussed in terms of the pertinent literature. While KC was independent of AgNP addition, porosity and grain size, both E and H decreased linearly with increasing porosity. In addition, this study is the first to identify (i) the Ag3Sb phase formed and (ii) the enhanced densification that occurs when the AgNP is sintered with Ba0.3Co4Sb12 powders, where both effects are consistent with the eutectic and peritectic reactions observed in the binary phase diagram Ag-Sb. These eutectic/peritectic 133 reactions may also be linked to the enhancement of electrical conductivity previously observed when Ag is added to Ba0.3Co4Sb12. Also, similar beneficial eutectic/peritectic reactions may be available for other systems where conductive particles are added to other antimonides or other thermoelectric systems. Keywords: thermoelectric; skutterudite, porosity, elastic modulus, peritectic reaction 6.1 Introduction The conversion of waste heat to electrical energy by thermoelectric (TE) materials is typically characterized by the material’s figure of merit, ZT, such that (1) where S is the Seebeck coefficient, σ is the electrical conductivity, κ is the thermal conductivity, and T is temperature [Ioffe 1960]. ZT values of up to 1.7 [Shi 2011] have been reported for bulk skutterudite TE materials. In-service conditions subject thermoelectric materials to stresses from thermal gradients, thermal transients, thermal expansion mismatch, and applied external stresses. Due to these challenges, the mechanical properties of thermoelectric materials are important, especially in waste heat harvesting applications. The response of a material to imposed stresses requires knowledge of the elastic properties; for example the Young’s modulus and Poisson’s ratio are needed in order to construct the stiffness matrix for finite element analysis [Segerland 1984; Hutton 2004; Zienkiewicz 2005]. Also, the fracture toughness, KC, which measures the resistance to crack growth [Wachtman 2009], is 134 important to the mechanical integrity. In addition, hardness is related to wear characteristics of the material [Ren 2008]. A key motivation for this study is the recent work on Ba0.3Co4Sb12 by Zhou et al. [Zhou 2012] which showed the addition of 0.5 wt% silver nanoparticles (AgNP) lead to a 30% increase in ZT, which was attributed primarily to an increase in electrical conductivity. Also, while carrier concentrations remain unchanged, a modest increase in the Seebeck coefficient also was observed with the added AgNP [Zhou 2012]. In general, the nanostructural and microstructural details can be extremely important for the transport properties of thermoelectric materials, and many previous studies have used nanostructures to reduce thermal conductivity by enhancing phonon scattering [Androulakis 2007; Zhao 2012; Zhou 2008; Alleno 2009; Ji 2007; Toprak 2004; Mi 2008], which can lead, in turn, to increases in ZT. Also, reducing the grain size or increasing porosity can enhance phonon scattering and thus increase ZT [Yoon 2013; Tokiai 1997]. Although in this study, an increased porosity leads to decreases in the elastic moduli, an increase in porosity also has the potential to increase the resistance to thermal fatigue damage [Case 2012b; Case 2012a], which is important due to the inevitable thermal cycling that thermoelectrics are subjected to during waste heat harvesting. In addition to affecting mechanical properties, porosity also impacts transport properties. In thermoelectrics such as nanoporous GeSi alloys [Lee 2010], porosity decreases both the electrical and thermal conductivity, although the decrease in electrical conductivity overwhelms the decrease in thermal conductivity, resulting in a net lowering of ZT due to porosity. However, in skutterudites including La0.75Fe3CoSb12 [Yang 2004], 135 Co0.9Ni0.1Sb3 , Co0.8Ni0.2Sb3 [He 2008] and a TiN nanoparticles-Co4Sb11.5Te0.5 composite [Wen 2013], enhancements in ZT have been reported for porosity levels between P = 0.003 and 0.15. For example, Wen et al. [Wen 2013] densified a 0.01 volume fraction TiN nanoparticles -Co4Sb11.5Te0.5 skutterudite and subsequent thermal annealing at 773 K for 120 hours in vacuum led to bloating (the formation of internal porosity) and a 10 percent increase in porosity which Wen et al. ascribed to a reaction between the TiN nanoparticles and the Co4Sb11.5Te0.5 matrix phases. Wen et al. noted that “The electrical conductivity and the thermal conductivity of the annealed sample decline simultaneously due to the higher porosity, but the thermal conductivity reduces more remarkably”, leading to a 20 percent increase in ZT compared to the denser specimens (prior to annealing and bloating) [Wen 2013]. Thus, in addition to having the potential to boost thermal fatigue resistance, in some skutterudites porosity can increase ZT making it important to also assess the mechanical properties such as elastic moduli, hardness and toughness for porous thermoelectrics. In this study, we examine the Young’s modulus, E, shear modulus, G, bulk modulus, B, Poisson’s ratio, , hardness, H, and fracture toughness, KC, of Ba0.3Co4Sb12 with and without the addition 0.5 wt% AgNP. We further explore the effects of microstructure/nanostructure and processing pursuing two processing routes: (i) planetary milling (PM), then hot press (HP) sintering and (ii) vibratory milling (VM) followed by pulsed electric current sintering (PECS). In this way, we produced specimens with a range of porosity and grain size while using the two most common milling techniques (planetary milling and vibratory milling) and densification techniques (HP and PECS) used to process thermoelectric materials. 136 6.2 Experimental procedure 6.2.1 Materials and Specimen Preparation To fabricate Ba0.3Co4Sb12, elemental Ba (pieces, 99.9% pure), Co (powder 99.5% pure), and Sb (shot 99.999% pure) in stoichiometric ratio were placed inside evacuated carbon coated quartz ampoules, heated to 1100°C and held at that temperature 5 hours, then quenched into a supersaturated salt water solution. The resulting ingots were then annealed at 750°C for 7 days. Finally, the ingots were planetary ball milled. In this study, two different processing techniques were used to fabricate two sets of specimens with differing microstructures. First, AgNP powders, which had a vendor specified purity of 99.9% and a size range of from 20 nm to 40 nm (45509, Alfa Aesar, Ward Hill MA), were dispersed into the Ba0.3Co4Sb12 powder by planetary ball milling at 300 rpm for 15 min [Pilchak 2007]. Then specimens approximately 10.5 mm in diameter and 10 mm in height were fabricated by hot pressing Ba0.3Co4Sb12 powders (with and without the added AgNP) in a graphite die for 20 minutes in an argon atmosphere with 50 MPa pressure at temperatures of 673 K, 773 K or 873 K in order to produce specimens with a range of volume fraction porosity, P, from 0.30 to 0.05 (Table 6.1). The hot pressed specimens were then cut into discs approximately 1.7 mm thick using a slow speed saw (Isomet, Buehler, Lake Bluff IL) with 0.3 mm thick diamond blade (801-137, Leco Corp., St. Joseph, MI). A second set of specimens was fabricated by reprocessing the specimens hot pressed without AgNP additions by first grinding the hot pressed specimens without Ag addition in an automated mortar and pestle (PM100, Retch) and then sieving with a 75 µm sieve, where the grinding and sieving steps were performed in an argon-filled glove 137 box containing less than 10 ppm oxygen. Using Viton gaskets (WC mill jar set 8004, WC media 5004A, Viton gasket 39322, SPEX Sample Prep, Metuchen, NJ) the resulting 7.1 g ground powder batches were sealed in a WC mill jar with a total of six spherical WC milling media (two WC spheres 11.2 mm in diameter and four WC spheres 7.9 mm in diameter). In order to perform vibratory milling outside the glove box, prior to removing the mill jar from the glove box, the mill jar and end cap were wrapped with electrical tape (Super 88, 3m, St. Paul, MN) and Parafilm “M” (PM-996 Pechiney Plastic Packaging, Menasha WI). The powders were milled for 10 minutes in a vibratory mill (8000M SPEX SamplePrep) then the mill jar was returned to the glove box. The reading of the oxygen meter on the glove box did not change when the mill jar was opened in the glove box. The oxygen meter is sensitive to changes within ± 0.1 ppm, implying that the sealing of the mill jar was effective. The vibratory milled powder then was sintered in an argon atmosphere by PECS (SPS Model 10-3, Thermal Technology LLC, Santa Rosa, CA) in a grafoil-lined graphite die at 673 K, 723 K or 773 K for 10 minutes with 50 MPa pressure (Table 6.1), producing disk-shaped specimens approximately 12.6 mm in diameter and 2.6 mm thick. 6.2.2 Elasticity Measurements Resonant ultrasound spectroscopy (RUS) elasticity measurements were performed using a tripod arrangement of piezoelectric transducers. A driving transducer which was swept through a range of frequency from 35 kHz to 485 kHz excited mechanical resonances in the specimen which were detected via the two pickup transducers. The peaks in the intensity versus frequency output of the pickup transducers (Figure 6.1) 138 Table 6.1. Specimens of skutterudite Ba0.3Co4Sb12 used in this study varied based on AgNP addition, sintering procedure and temperature, resulting porosity, and powder processing. Specimens VM-PECS-673, VM-PECS-723 and VM-PECS-773 were reprocessed from specimens PM-HP-673, PM-HP-773 and PM-HP-873 by grinding and powder processing by vibratory milling, and sintering by PECS. Specimen AgNP Powder addition processing Sintering procedure Sintering temperature PM-HP-673 No HP 673K Vol. fraction porosity 0.30 773K 873K 673K 773K 873K 673K 723K 773K 0.21 0.16 0.27 0.18 0.05 0.16 0.13 0.09 PM PM-HP-773 No PM HP PM-HP-873 No PM HP PM-HP-673-Ag Yes PM HP PM-HP-773-Ag Yes PM HP PM-HP-873-Ag Yes PM HP VM-PECS-673 No VM PECS VM-PECS-723 No VM PECS VM-PECS-773 No VM PECS PM: Planetary milling, VM: Vibratory milling, HP: Hot pressed, PECS: Pulsed electric current sintering Intensity (a.u.) 45 30 15 0 0 200 400 600 Frequency (KHz) Figure 6.1. RUS spectrum from specimen PM-HP-873-Ag, P = 0.05 with AgNP. The elastic moduli of each of the specimens in this study are calculated from the specimen mass, dimensions, shape, and resonant frequencies. Each peak in the spectrum represents a mechanical resonance at that frequency 139 represent the mechanical resonance frequencies of the specimen. From the resonant frequencies, specimen mass, dimensions and geometry, the elastic moduli of the disk spaced specimens were calculated using commercial software (RUSpec and CylModel, Quasar International, Inc., Albuquerque, NM). Additional details of the RUS procedure are provided elsewhere [Ren 2009; Ni 2010; Schmidt 2010; Migliori 1997; Schmidt 2013]. 6.2.3 Hardness and toughness measurements Prior to indentation measurements of H and KC, and prior to energy-dispersive Xray spectroscopy measurements, the specimens were polished with diamond paste with grit sizes ranging from 30 µm to 1 µm. At least ten indentations per load for loads of 1.96 N, 2.94 N, 4.9 N and 9.8 N were made on the polished surface of each specimen, with a loading time of 5 seconds for each indentation. Prior to the hardness measurements, the Vickers indenter was calibrated using a steel standard calibration block (Yamamoto Scientific Tools Lab Co. LTD, Chiba, Japan). The hardness calibration factor, ζranged from 0.95 to 0.97 for the indentation loads used in this study. The hardness, H, for indentation was calculated from [Lawn 2012] H= 1.8544 F (2) ( 2 a )2 where F is the applied indentation load and a is half of the diagonal length of the indentation impression. The fracture toughness, KC, was calculated from KC   ( E / H ) 0.5 F (3) c3/ 2 140 where ξ is a dimensionless calibration constant equal to 0.016 [Wachtman 2009], E is the Young’s modulus measured by RUS, H is the hardness value and c is half of the radial crack length and F is the applied load [Wachtman 2009] . 6.2.4 Energy-dispersive X-ray spectroscopy and microscopy Except for specimen PM-HP-673-Ag, each of the sintered specimens was first polished and then examined by energy-dispersive X-ray spectroscopy (EDS). However, polishing was unsuccessful for PM-HP-673-Ag, so the examined surfaces were those cut by a low speed diamond saw. The microstructure and nanostructure of the powders and sintered specimens was observed by scanning electron microscope (SEM, JEOL 6610LV or JSM-7500F, JEOL Ltd., Japan) at a 15 kV accelerating voltage and either a 15 mm or 8 mm working distance. Using SEM micrographs of the fracture surfaces of sintered specimens, the grain size of sintered specimens was determined using the linear intercept method with at least 200 intercepts per micrograph and a stereographic projection factor of 1.5 [Underwood 1969; Case 1981]. The microstructure of the polished specimen surfaces was observed using the secondary electron image (SEI) mode. Atomic number contrast was observed in the backscattered electron image (BEI) mode. Compositions were determined in the BEI mode using EDS. For SEM observation, all specimens were mounted on aluminum stubs, where the silver nanoparticles were adhered to the stubs using carbon paint (05006-AB, SPI Supplies, West Chester, PA) and the sintered specimens were mounted using carbon tape. 141 Figure 6.2. Silver nanoparticles exhibited agglomerates of 10 µm or greater (a-c), but consisting of individual grains or particles consistent with the manufacturer claimed average particle size of 20-40 nm (d) 142 Figure 6.3. Polished surfaces of specimen PM-HP-873-Ag, both in secondary electron images (a) and backscatter images (b). Porosity is observed between larger grains (a and b), but also with areas of Ag visible in backscatter as bright areas (b), due to the higher average atomic weight of the silver rich regions. The abundance of Ag and the relative deficiency of Co or Sb in the bright areas of the backscatter image (b) is confirmed by energy-dispersive x-ray spectroscopy maps (c-e) 143 Counts per Second 300 (b) Sb Co Ag PM-HP-673-Ag Ag 200 Sb 100 Co 0 0 10 20 30 Counts per Second Distance (µm) Sb Co Ag (d) PM-HP-773-Ag 600 Ag Sb 400 200 Co 0 0 1 2 3 4 Distance (µm) Sb Co Ag Counts per Second (f) PM-HP-873-Ag 600 Ag Sb 400 200 Co 0 0 1 2 3 4 5 Distance (µm) Figure 6.4. For each of the hot pressed specimens with AgNP additions, an EDS line scan was performed and point ID locations were chosen. For PM-HP-673-Ag, a cut surface was used (a) because the specimen was not successfully polished. For PM-HP-773-Ag and PM-HP-873-Ag, polished surfaces were examined (c and e). The line scan for PMHP-673-Ag (b) indicated only silver present except for two regions in the silver where antimony and cobalt were present in concentrations consistent with skutterudite particles. In contrast, the line scan for PM-HP-773-Ag (d) and PM-HP-873-Ag (f) indicated the presence of antimony in the silver-rich locations. No silver was observed in the matrix outside of the silver-rich locations 144 Figure 6.5. The microstructure of Ba0.3Co4Sb12, without Ag addition, changed as a function of sintering temperature. For (a) a sintering temperature of 673 K, microstructure and porosity, P, are consistent with a specimen with little to no observed sintering or densification. After sintering at 773 K (b), only minimal densification and neck formation are observed. Hot pressing at 873 K (c) enhanced both grain growth and densification in the specimen without Ag nanoparticles, however, the microstructure still is observed to be very porous, consistent with measurement of P = 0.16 145 Figure 6.6. The microstructure of Ba0.3Co4Sb12 with 0.5 wt% AgNP addition changed as a function of sintering temperature. At 673 K (a), little to no sintering is observed in the microstructure. Hot pressing at 773 K (b) resulted in limited grain growth and densification, with necks observed between the grains but significant porosity observed, consistent with early stage sintering. The porosity for the specimen with Ag nanoparticles sintered at 873 K (c) is the lowest (P = 0.03) of all the specimens in this study 146 Figure 6.7. Fracture surface images of reprocessed material sintered by PECS exhibit dense regions surrounded by regions of higher porosity. All regions of the specimens exhibit similar grain sizes. The porous areas decrease as the sintering temperature increased from (a) 673 K, to (b) 723 K, to 773 K 147 Figure 6.8. After reprocessing by SPEX milling and sintering by PECS, the samples exhibit a unimodal, sub-micron grain size distribution. 148 6.3 Results and Discussion 6.3.1 Microstructural and chemical analysis The addition of AgNP resulted in a greater change in density at the higher sintering temperature than the specimen without AgNP addition (Table 6.1). When HP sintered at 673 K, the porosity was nearly identical for the specimen without (P = 0.30) and with AgNP (P = 0.27), but when sintered at 873 K, the specimen without AgNP addition (P = 0.16) was significantly less dense than the specimen with AgNP addition (P = 0.05). The density difference for specimens sintered at the same temperature requires a close examination of the microstructural features of the AgNP addition, examined in two parts. The microstructural and nanostructural features examined in this study included: (i) the as-received AgNP powders (Figure 6.2), and for the sintered Ba0.3Co4Sb12 specimens the micron-sized Ag inclusions (Figures 6.3 and 6.4) in section 6.3.1.1, and (ii) porosity (Figures 6.5 – 6.7) as well as grain size and size distribution (Figure 6.5 – 6.8) in section 6.3.1.2. 6.3.1.1 Microstructure of the AgNP and Ag agglomerates The as-received AgNP powders consisted of (i) large, irregularly-shaped agglomerates up to 30 µm across, with most agglomerates being between 0.5 µm and 10 µm in diameter, (ii) individual spherical nanoparticles with diameters 30 nm (Figure 6.2d) and (iii) small clusters of nanoparticles (Figure 6.2). Similar agglomerates, small clusters and individual AgNP are visible in micrographs taken by Zhou (Figure 2a of Zhou 2012]) of AgNP powders received from the same vendor as in this study. After hot pressing at 673 K, 773 K and 873 K, the consolidated, micron-size Ag particles observed in the Ba0.3Co4Sb12 specimens with 0.5 wt% AgNP (Figures 6.3 and 149 6.4) had roughly the same dimensions as the original AgNP agglomerates present in the as-received AgNP powders (Figure 6.2). In the sintered specimens, the typical dimension of Ag particles observed in PM-HP-773-Ag and PM-HP-873-Ag was between 0.3 µm and 10 µm, with a few Ag particles of 20 µm diameter or greater, as observed by backscatter electron images of polished surfaces (Figures 6.3 and 6.4). Thus the AgNP agglomerates may not have broken up significantly during mixing in the planetary mill prior to hot pressing. As a rough gauge of how easily the AgNP agglomerates can be broken apart by milling, 0.0225g of AgNP was wet milled by hand for 60 sec in a porcelain mortar and pestle (60310 and 60311, CoorsTek, Golden, CO) with 5 mL of ethanol. After milling, the AgNP agglomerates were still present but some surfaces of the agglomerate surfaces appeared to be deformed by the hand milling and some of the nanoparticles on the deformed surfaces of the agglomerates appeared to be smeared (Figure 6.9). After hand milling, the AgNP agglomerates typically ranged in diameter from sub-micron to 100 µm (Figure 6.9), similar to or greater than the dimensions of AgNP agglomerates before milling. Thus, based on both the SEM examination of the AgNP agglomerates before and after hand milling, the hand milling process did not appear to significantly break up the agglomerates. While the hand milling conditions are likely different than those encountered during the planetary milling, it is likely that some of the original AgNP agglomerates in the as-received powders survived (at least partially intact) the mixing process with the AgNP agglomerates powders to give the micron-sized consolidated Ag particles observed in the as-sintered Ba-skuttterudite-AgNP specimens in this study. This is further 150 Figure 6.9. Hand milling of AgNP agglomerates in ethanol were not observed to reduce the size of the nanoparticle agglomerates. After hand milling, silver nanoparticle agglomerates of 10 µm or greater were typically observed, similar to the size of agglomerates for the as received AgNP. Several faces of the agglomerates appear to be deformed after hand milling 151 supported by the SEM-based observation that the size of the AgNP agglomerates in the powders is similar to the size of the Ag particles in the sintered specimens. Thus, it does not seem to be necessary to postulate an accretion of AgNP during milling to explain the present of micron-sized Ag particles in the sintered specimens. 6.3.1.2 Porosity and grain size and relationship to the processing technique The six hot pressed specimens included this study (Table 6.1) showed a bimodal grain size distribution (Figures 6.5 and 6.6) consisting of a polycrystalline matrix with grain sizes ranging from approximately 0.3 µm to 3 µm, with an average grain size (GS) of roughly 1 µm. Also, grains approximately 5 µm to 20 µm across were observed in each HP specimen, both without AgNP (Figure 6.5) and with AgNP (Figure 6.6) additions. Pores with diameters of submicron to a few microns across were distributed through the matrix, consisting of open, interconnected porosity when sintered at 673 K, and consisting of partially to fully isolated porosity at 773 K and 873 K (Figures 6.5 and 6.6). The observed pore morphologies for the specimens sintered at 673 K, 773 K and 873 K agrees with the pore evolution in sintered powders [Barsoum 2003]. Each of the three reprocessed specimens (vibratory milled and then sintered via PECS at 773 K, 723 K and 673 K) had a much different microstructure than those obtained by planetary milling and hot pressing. Namely, the VM-PECS specimens a unimodal GS distribution (Figure 6.8) with a significantly smaller average GS of 0.17 µm, 0.14 µm, and 0.15 µm, respectively, than the specimens that were planetary milled and hot pressed. The reprocessed specimen VM-PECS-773 had a volume fraction porosity of only 0.09, which included dense regions with clusters of pores (Figure 6.7c). However, specimen VM-PECS-673 included clusters of pores with islands of denser 152 material (Table 6.1, Figures 6.7a). The microstructure of specimen VM-PECS-723 was intermediate in between specimens VM-PECS-773 and VM-PECS-673 (Table 6.1, Figures 6.7b). The addition of AgNP results in an increased density of hot pressed specimens at a sintering temperature of 873 K, where the specimen without AgNP had a P value of 0.16 while the specimen sintered with 0.5 wt% AgNP had P = 0.05 (Table 6.1). At a hot press sintering temperature of 773 K, the porosities of the specimens with and without 0.5 wt% AgNP were 0.18 and 0.21, respectively (Table 6.1). Also, for a sintering temperature of 673 K, the porosities of the specimens with and without 0.5 wt% AgNP were 0.27 and 0.30, respectively (Table 6.1). Thus, hot press sintering at 773 K and 673 K resulted in similar porosities for specimens with and without 0.5 wt% AgNP additions. However, the AgNP addition resulted in considerably enhanced densification for hot press sintering at 873 K. The enhanced sintering is likely related to a reaction between the 0.5 wt% AgNP addition and the Ba0.3Co4Sb12 matrix, as discussed in section 6.3.2. 6.3.2 Chemical analysis 6.3.2.1 Chemistry of the AgNP and Ag agglomerates EDS analysis showed the as-received AgNP powder to be 99.5 at% Ag, with oxygen and Al impurities (Table 6.2). The Al in the EDS analysis likely originated from the aluminum stub on which the specimens were mounted. Polished surfaces of specimens PM-HP-873-Ag and PM-HP-773-Ag were examined by EDS, with matrix region and Ag-rich regions in each specimen examined separately at the points indicated in Figures 6.4c and 6.4e, and line scans extending 153 Table 6.2 Results of EDS scan using Point ID mode on two as-received AgNP agglomerates. Location Ag Al O Ag:Al label AgNP-1 85.15 0.60 14.25 99.3:0.7 AgNP-2 99.5 0.5 Not 99.5:0.5 Detected Table 6.3 Results of EDS scan using Point ID mode on specimen PM-HP-873-Ag. Spacing between spots is approximately 5 to 10 µm (Figure 6.4e). Ag areas examined are 1-3 µm in diameter. Location Ba Co Ag Sb O label Matrix 1.69 21.54 Not Detected 76.78 Not Detected Ag-873-1 Not Detected 1.28 73.96 24.76 Not Detected Ag-873-2 Not Detected Not Detected 76.78 23.22 Not Detected Ag-873-3 Not Detected 1.37 71.82 22.54 6.78 Table 6.4 Results of EDS scan using Point ID mode on specimen PM-HP-773-Ag. Spacing between spots is approximately 4 µm (Figure 6.4c). Ag area examined approximately 1-3 µm in diameter. Location Ba Co Ag Sb O label Matrix 2.14 23.96 Not Detected 73.89 Not Detected Ag-773-1 Not Detected 2.49 70.01 27.50 Not Detected Table 6.5 Results of EDS scan using Point ID mode on specimen PM-HP-673-Ag. Spacing between spots is approximately 5 to 10 µm (Figure 6.4a). Ag area examined is approximately 20 µm in diameter. Location Ba Co Ag Sb O label Matrix 1.44 21.01 Not Detected 77.56 Not Detected Ag-673-1 Not Detected 6.80 72.21 20.99 Not Detected Ag-673-2 Not Detected Not Detected 92.99 7.01 Not Detected Ag-673-3 Not Detected Not Detected 94.21 5.79 Not Detected 154 across both matrix material and Ag-rich regions (Figure 6.4c – 6.4f). As determined by EDS, the elemental composition of the matrix region is consistent with Ba0.3Co4Sb12 (Tables 6.3 and 6.4, Figure 6.4c – 6.4f). In the Ag-rich regions, the composition is consistent with primarily Ag3Sb phase material. In specimen PM-HP-873-Ag, the Ag:Sb ratio was 3.0:1 to 3.3:1 in the silver-rich regions (Table 6.3, Figure 6.4e), and in specimen PM-HP-773-Ag, the ratio was 2.5:1 (Table 6.4, Figure 6.4c). Less than 2 at% Co was detected in each of the silver-rich regions, consistent with little to no skutterudite phase present in the silver-rich regions. In contrast to the preparation of specimens PM-HP-873-Ag and PM-HP-773-Ag, specimen PM-HP-673-Ag was cut on a low speed diamond saw because the high porosity of the specimen (P = 0.27, Table 6.1) made it too friable for polishing. The elemental composition of the matrix region is consistent with Ba0.3Co4Sb12 (Table 6.5, Figure 6.4a and 6.4b). In the Ag-rich regions of PM-HP-673-Ag, the Ag:Sb ratio was highly variable, with ratios of 3.4:1, 13.3:1 and 16.3:1 (Table 6.5), which is more indicative of a mixture than a single phase. In the Ag-rich location with the highest concentration of Sb, Ag-673-1, a significant concentration of Co was also detected, 6.80 at%, and in a Co:Sb ratio of 1:3.1 (Table 6.5), consistent with a particle or particles of Ba0.3CoSb3 (matrix material) embedded in otherwise pure Ag. Particles of Ba0.3CoSb3 may have become embedded in the Ag-rich region during cutting. The presence of small, discrete skutterudite particles embedded in the silver in specimen PM-HP-673-Ag is further supported by the line scan (Figure 6.4b), in which there are two dips in the counts of silver at distances of 5 µm and 10 µm, with corresponding increases of both Co and Sb counts at each location. The EDS results from the other two Ag-rich locations, Ag-673-2 155 and Ag-673-3 (Table 6.5), are primarily pure Ag material, with no indication of embedded Ba0.3CoSb3 particles. 6.3.2.2 Sintering behavior changes and scavenging of Sb Although the Ba0.3Co4Sb12-Ag composition is within a quaternary Ba-Co-Sb-Ag system for which no phase diagram is available, the Ag-Sb binary phase diagram may provide important guidance to understanding the formation of Ag3Sb particles. In the binary phase diagram for Ag-Sb, there are only two intermediate compounds, Ag3Sb and Ag6Sb [Voronin 2013; Cipriani 1996; Feschotte 1992]. Upon cooling, Ag3Sb forms via a peritectic reaction Ag6Sb + L  Ag3Sb at 835 + 2 K and an initial Sb concentration of 21.2 at% [Feschotte 1992]. Also, the eutectic reaction L  Ag3Sb + Sb occurs at 757 K with a eutectic composition reported as 38 at% Sb [Voronin 2013; Okamoto 2007; Hassam 2001] or 41 at% Sb [Okamoto 1993]. In the present research, the Sb that reacts with the AgNP may be scavenged from the Ba0.3Co4Sb12 matrix. Scavenging of excess Sb can aid thermal stability in skutterudites. For example, Zhang and Sakamoto found that the Sb deficient specimens of Fe3.5Co0.5Sb12-based skutterudites were more dimensionally stable than specimens with excess Sb [Zhang 2013]. The presence of a transient liquid phase is known to greatly enhance densification during sintering [Barsoum 2003], thus the enhanced densification for the Ba-skutteruditeAgNP specimens observed in this study (section 3.1, Table 6.1) may be due to a transient liquid phase during sintering associated with the peritectic and eutectic reactions in the Ag-Sb system. Also, if the liquid phase wets the grain boundaries, the electrical conductivity of the grain boundaries could be greatly enhanced, by a surface film rather 156 than only point contacts due to discrete AgNP or Ag particles, as was suggested by Zhao et al. [Zhou 2012]. A similar binary and ternary eutectic in lead-free Sn-Ag-Cu solders aids in wetting the surfaces of electrical wires, pins or pads [Bukat 2011; Dharma 2009]. 6.3.2.3 Possible significance to other thermoelectric material systems The proposed reaction of Ag with Sb in Ba-filled skutterudite Ba0.3Co4Sb12 has important implications for other skutterudites, other antimonides and even other thermoelectric systems. In this study the Ba-filled skutterudite is CoSb3-based, but more generally unfilled skutterudites have the general composition TX3, where T is a transition metal element (Co, Fe, Ni) and X is a pnicogen element (Sb, As, and P) [Yoshizawa 2004]. “Filler” atoms, often a rare earth element or mischmetal added to the skutterudite composition, scatter phonons to make them an important class of thermoelectric materials [Slack 1994; Morelli 1995; Nolas 1999; Nolas 2000; Uher 2001]. Many CoSb3-based and FeSb3-based compositions have been studied in the literature [Zhang 2010], so in addition to the particular Ba-filled skutterudite included in this study, adding AgNP to other antimonide skutterudites may well lead to the same type of Ag-Sb reaction and thus liquid phase sintering, the formation of Ag3Sb, enhanced densification and improvements in the ZT and electrical conductivity. Also, for an antimonide thermoelectric that is not a skutterudite, a recent study by [Xiong 2013] shows that Ag added to ZnSb leads to the formation of Ag3Sb particles and a higher electrical conductivity. This type of reaction may not be limited to Ag and antimonides. In fact, there may be other conductive metal-thermoelectric material combinations in which similar peritectic/eutectic reactions occur, accompanied by similar beneficial effects. One guide to searching for such systems would be appropriate binary or ternary phase diagrams. 157 6.3.3 Elasticity results The measured elastic moduli are essentially independent of the AgNP addition and the processing/grain size, but are strong functions of porosity (Figure 6.10a-c). 6.3.3.1 Elasticity as a function of AgNP addition In order to determine whether or not the elastic moduli are functions of the AgNP addition, for the set of specimens (with and without AgNP addition) included in this study, we examined the porosity dependence of Young’s modulus, shear modulus and bulk modulus (Table 6.1, Figure 6.10). There are no significant differences between the slope and intercept for the least squares fit to the data (i) with and (ii) without AgNP additions, with coefficient of determination, R2, values of 0.955 to 0.998 for E and G, and R2 of 0.857 and 0.947 for B (Figure 6.10). To put the elasticity dependence on AgNP addition into context, the literature shows that the elastic modulus as a function of the overall composition depends on the physical nature of the material system itself, where compositional changes can occur when one has (i) a composite system in which discrete particles are added to a matrix of differing composition or (ii) a solid solution system or (iii) a solid solution system that includes micro- or nano scale particles. Changes in elastic moduli of a particulate composite as a function of the addition of a given volume percent of particulate phase has been modeled by a number of researchers [Hashin 1962; Halpin 1992; Bedolla 2012; Couturier 1997] (Appendix A). The four composite models given in Appendix A predict a decrease in the elastic modulus of the composite, EC of about 0.39% with the 0.5 wt% AgNP addition to the Ba-skutterudite in this study. 158 Without AgNP 150 (a) Without AgNP (b) With AgNP Without AgNP, reprocessed 60 G (GPa) E (GPa) 120 90 60 With AgNP Without AgNP, reprocessed 30 30 0 0.0 0.1 0.2 0 0.0 0.3 0.1 Porosity 0.2 0.3 Porosity Without AgNP 90 (c) 0.25 With AgNP Poisson's Ratio Without AgNP, reprocessed B (GPa) 60 30 0 0.0 0.1 0.2 (d) 0.20 0.15 0.10 Without AgNP 0.05 0.00 0.0 0.3 Porosity With AgNP Without AgNP, reprocessed 0.1 0.2 0.3 Porosity Figure 6.10. The (a) Young’s modulus, E, (b) shear modulus, G, and (c) bulk modulus, B, are each a function of porosity. In each figure, the solid lines represent a least-squares fit to equation (5b). The Poisson’s ratio (d) is observed to be a weak function of porosity. The elastic moduli were not observed to be a function of the addition of 0.5 wt% AgNP or of reprocessing 159 Table 6.6. The linear decrease in elastic moduli, E, G, and B, according to equation (5b), for this study of Ba0.3Co4Sb12 is consistent with the limited information available in the literature for porosity dependence of elastic moduli for thermoelectric materials [Schmidt 2013; Ni 2009]. Material A0 (GPa) b N Prange R2 Reference Ba0.3Co4Sb12 + 0.5 143 ± 4 (E0) wt% AgNP, E Ba0.3Co4Sb12 , E 146 ± 9 (E0) YbAl3 , E 174.0 ± 2.5 (E0) LAST, E 58.3 ± 0.3 (E0) Ba0.3Co4Sb12 + 0.5 59 ± 2 (G0) wt% AgNP, G Ba0.3Co4Sb12 , G 60 ± 3 (G0) YbAl3, G 73.6 ± 0.9 (G0) LAST, G Ba0.3Co4Sb12 + 0.5 wt% AgNP, B Ba0.3Co4Sb12 , B YbAl3 , B 3.11 ± 0.08 3 0.05-0.27 0.998 This study 3.00 ± 0.16 2.34 ± 0.06 6 7 0.09-0.30 0.03-0.23 3.6 ± 0.1 3.07 ± 0.08 12 0.01-0.14 3 0.05-0.27 0.955 This study 0.994 [Schmidt 2013] 0.994 [Ni 2009] 0.998 This study 2.96 ± 0.15 2.38 ± 0.06 6 7 0.09-0.30 0.03-0.23 22.9 ± 0.1 (G0) 3.5 ± 0.1 82 ± 2 (B0) 3.25 ± 0.05 12 0.01-0.14 3 0.05-0.27 87 ± 10 (B0) 3.14 ± 0.31 91.3 ± 2.9 (B0) 2.22 ± 0.14 6 7 160 0.09-0.30 0.03-0.23 0.962 This study 0.995 [Schmidt 2013] 0.991 [Ni 2009] 0.999 This study 0.857 This study 0.947 [Schmidt 2013] 6.3.3.2 Elastic moduli and porosity The elastic moduli of the skutterudite Ba0.3Co4Sb12 in this study are linear functions of porosity over the range of volume fraction porosity included in this study (between 0.05 and 0.30, Table 6.6, Figure 6.10). In addition, the elastic modulus versus porosity data is independent of either the presence or absence of the 0.5 wt% AgNP (Table 6.6, Figure 6.10). Also, the powder processing technique (planetary milling or vibratory milling) and the densification method (hot pressing or PECS (Table 6.1)) does not affect the elastic modulus versus porosity behavior (Table 6.6, Figure 6.10). The elastic moduli of TE materials are porosity dependent [Schmidt 2013; Ni 2009]. An empirical expression frequently used to describe the porosity dependence of Young’s modulus, E, shear modulus, G, and bulk modulus, B [Schmidt 2013; Ni 2009] of brittle materials [Rice 1998], including TE materials [Schmidt 2013; Ni 2009] A = A0(exp-bAP) (5a) where A represents the property E, G, or B [Rice 1998] and bA is a unitless, materialdependent parameter that measures the rate of decrease in property A with increase P . If bAP is small, the equation may be linearized by using the first two terms of the Taylor series expansion of equation (5a) [Schmidt 2013]. A = A0(1 - bAP) (5b) For the nine Ba0.3Co4Sb12 specimens in this study (with and without AgNP addition), a least-squares fit of the modulus data to equation (5b) yielded P = 0 intercept values of E0 = 145.1 ± 5.2 GPa, G0 = 59.8 ± 2.0 GPa, and B0 = 84.9 ± 5.9 GPa with coefficients of determination, R2, of at least 0.90 for E, G, and B. There is no significant difference between the P = 0 intercept for the three specimens with 0.5 wt% AgNP as 161 compared to the six specimens without AgNP addition (Table 6.6). Also, the rate of decrease in modulus with porosity (given by bA in equation (5b)) was between 3.0 and 3.2 for the Ba0.3Co4Sb12 (Table 6.6), which is consistent with values of 2 to 6 for a wide range of solid materials [Rice 1998]. Although the data for the porosity dependence of elastic moduli of thermoelectric materials is relatively limited, modulus versus P data is available for YbAl3 [Schmidt 2013], LAST [Ni 2009] (Table 6.6) and several for Co4Sb12-based skutterudites other than the Ba-filled skutterudite included in this study (Figure 6.11) [Zhang 2010]. The bA values for the Ba-filled skutterudite for E, G, and B in this study are bracketed by the bA values for YbAl3 and LAST for the corresponding moduli (Table 6.6). Also, the elastic modulus-porosity behavior for Ba0.3Co4Sb12 in this study is quite similar to that reported by Zhang [Zhang 2010] for other Co4Sb12-based skutterudites (Figure 6.11). In this study, Poisson’s ratio is insensitive to P for porosities P < approximately 0.20, but there is considerable scatter in the data (Figure 6.10d). For two previous studies of TE materials, the Poisson’s ratio for LAST [Ni 2009] and YbAl3 [Schmidt 2013] was essentially constant or weakly decreasing with increasing P, for P less than 0.15 or 0.12. In this study, the scatter in Poisson’s ratio (Figure 6.10d) is typical of observations for porous ceramics [Ramakrishnan 1993; Boccaccini 1994], and the observed relative decrease in Poisson’s ratio at P greater than about 0.20 may be reasonably expected. Ramakrishnan and Arunachalam developed a model using 2dimensional finite element simulation showing Poisson’s ratio approaches 0.25 as porosity increases [Ramakrishnan 1993]. However, the model does not agree with the 162 Elastic Moduli (GPa) Ba0.3Co3Sb12 (this study) Ba0.075Sr0.025Yb0.1Co4Sb12 [Zhang 2010] 150 120 90 (BaYb)0.03Co4Sb12 [Zhang 2010] E Ca0.07Ba0.23Co3.95Ni0.05Sb12 [Zhang 2010] B 60 30 G 0 0.0 0.1 0.2 0.3 Porosity Figure 6.11. The elastic moduli for the exact composition of skutterudite in this study are not recorded in literature, however the moduli for specimens of similar composition [Zhang 2010] are consistent with the porosity dependent elastic moduli relationships observed in this study. Filled symbols indicate specimens with 0.5 wt% AgNP, unfilled for specimens without any AgNP, and half-filled symbols for reprocessed specimens without any AgNP. The solid lines represent a least squares fit to equation (5b) of the data in this study 163 available data on porous ceramics [Boccaccini 1994], and may be limited by the nature of the model, ignoring the heterogeneity of porosity [Rice 1995]. Furthermore, Rice argues that any model should converge to 0 for P = 1, and argues that Poisson’s ratio should generally decrease when approaching higher values of P [Rice 1995]. In an examination of both pore shape and P, Dunn and Ledbetter determine that Poisson’s ratio may increase, decrease or remain unchanged, depending on pore shape and distribution [Dunn 2011]. Thus for a Poisson’s ratio in the range of about 0.20 to 0.25 (for the pore-free material), the experimental observations [Ramakrishnan 1993; Boccaccini 1994; Rice 1995] and theoretical predictions [Ramakrishnan 1993; Dunn 2011] agree that Poisson’s ratio is nearly independent of P (for spherical pores) over the interval from approximately 0 < P < 0.20, which agrees with behavior observed for the Ba-skutterudites in this study (Figure 6.10d). 6.3.3.3 Elastic moduli and grain size/processing effects There was no observed difference between porosity dependent modulus behavior of the reprocessed specimens and the HP specimens (Figure 6.10). The elastic moduli of the skutterudite Ba0.3Co4Sb12 in this study (Figure 6.10) are not functions of grain size (Figures 6.5-6.8, Table 6.7) or processing (Table 6.1). Grain size varied from bimodal with grains up to approximately 20 µm across (Figures 6.5 and 6.6, Table 6.7) to less than 0.2 µm (Figures 6.7 and 6.8, Table 6.7), based on processing by PM and HP, or reprocessed by VM and PECS (Table 6.1). The bimodal grain size distribution is not expected to influence the elastic moduli of the material, as the elastic moduli of materials with grain sizes greater than approximately 20 nm are typically not a function of grain size [Kim 1999]. 164 Table 6.7. For CoSb3-based thermoelectric materials, a comparison of the KC results from this study with the literature, including the porosity, P, the grain size, GS, of the specimens and the KC measurement technique [Ravi 2009; Eilertsen 2013; Rogl 2011]. Material KC P GS KC Reference (MPam0.5) (µm) measurement method Ba0.3Co4Sb12 0.8 - 1.3 0.09 Bimodal a, Vickers This study 0.20 0.14-0.17 indentation Ba0.3Co4Sb12- 1.0 - 1.3 0.05 Bimodal a Vickers This study 0.5 wt% Ag 0.18 indentation CoSb3 1.7 < 0.01 NA Chevron notch [Ravi 2009] bend CeFe31.1 - 2.8 < 0.01 NA Chevron notch [Ravi 2009] bend xRuxSb12 CoSb3 0.82 ± Nearly 1-2 µm, Vickers [Eilertsen 0.11 dense 15 µm indentation 2013] CoSb3 0.52 ± Nearly 1-2 µm, Crack opening [Eilertsen 0.04 dense 15 µm displacement 2013] CoSb3 0.51 ± Nearly 15 µm Single-edge vee- [Eilertsen 0.06 dense notch bend 2013] 0.1In CoSb3 0.46 ± Nearly 15-40 µm Vickers [Eilertsen 0.13 dense indentation 2013] 0.1In CoSb3 0.49 ± Nearly 15-40 µm Crack opening [Eilertsen 0.03 dense displacement 2013] 0.1In CoSb3 0.57 ± Nearly 15 µm Single-edge vee- [Eilertsen 0.06 dense notch bend 2013] Fe4Sb12- and 1.5 to 2.2 Typically NA Vickers [Rogl 2011] b Fe3CoSb12< 0.02 indentation based b NA: Information not available a Bimodal specimens had matrix grain sizes of 0.3 µm to 3 µm, with larger grains approximately 5 µm to 20 µm across b DD0.88Fe4Sb12, P = 0.014; Ca0.41DD0.41Fe4Sb12, P = 0.015; Ba0.44DD0.42Fe4Sb12, P = 0.018; Ca0.20Sr0.12DD0.39Fe3CoSb12, P = 0.011; Ca0.20Ba0.14DD0.38Fe3CoSb12, P not given; Sr0.12Ba0.18DD0.39Fe3CoSb12, P = 0.010; Sr0.066Ba0.066Yb0.066Co4Sb12, P not given; Mm0.78Fe3CoSb12, P not given 165 The distribution of porosity may be influenced by the microstructure, with associated influences on elastic moduli, however previous studies of TE materials showed only a relation based on the volume fraction porosity and no pore distribution or shape factor required to describe the modulus-porosity relationship [Schmidt 2013; Ni 2009]. 6.3.4 Hardness and fracture toughness results Specimens PM-HP-673 and PM-HP-673-Ag, both hot pressed at 673 K, were too fragile to polish, limiting the hardness and fracture toughness measurements for the HP specimens to the two sintered at 773 K and the two sintered at 873 K. However, H and KC measurements were made on each of the three reprocessed PECS specimens since all of those specimens had sufficient mechanical integrity so that polishing and indentation was not a problem. Also, for the toughness measurements, radial cracks were not present at a Vickers indentation load of 1.96 N for any of the specimens except VM-PECS-673 and VM-PECS-773, nor at 2.94 N for specimen PM-HP-673, so no KC data is reported for the 1.96 N indentations for any specimen, nor at 2.94 N for specimen PM-HP-673 (Figure 6.12). 6.3.4.1 Hardness and toughness as a function of Ag addition As was the case for the elastic moduli, the hardness and fracture toughness of the Ba-skutterudite in this study were not sensitive to the AgNP addition (Figure 6.12, Table 6.7). For H, much of the literature on micro-sized and nanosized metallic particles added to brittle matrices has focused on higher volume percentages. For example, H of alumina decreased from 18 GPa to 9.5 GPa with the addition of 0.20 volume fraction aluminum 166 (a) Hardness (GPa) 6 9.8 N 4.9 N 2.94 N 1.96 N VM fit BM fit Vib ra tory mil led 4 Plan e 2 0 0.00 0.05 tary 0.10 mill ed 0.15 0.20 Porosity 1/2 Kc (MPa m ) 2.0 9.8 N 4.9 N 2.94 N Average KC (b) 1.5 StDev KC 1.0 0.5 0.0 0.00 0.05 0.10 0.15 0.20 Porosity Figure 6.12. The (a) hardness and (b) fracture toughness for the hot pressed specimens with AgNP, filled symbols, the hot pressed specimens without AgNP, open symbols, and the reprocessed PECS specimens without AgNP, half-filled symbols, from Vickers indentation at four loads, 9.8 N, 4.9 N, 2.94 N, and 1.96 N. Full radial cracks were not observed on the specimens at 1.96 N load and therefore no fracture toughness values are available at 1.96 N load. The hot pressed specimens sintered at 673 K were not able to be polished and were not tested. The solid lines in the figure (a) represent hardness for the reprocessed specimens were fit to equation (6) separately from the hot pressed specimens because hardness is a function of grain size. Fracture toughness, figure (b), was not observed to be a function of either porosity or load, as plotted by the average (solid) and standard deviation (dotted) lines, with an average KC for all 7 specimens of 1.0 ± 0.2 MPa-m½ 167 particles (average XRD crystallite size of 25 nm, TEM grain size 26 ± 3 nm) [Zawrah 2012], and from 8.2 GPa to 7.3 GPa in hydroxyapatite with the addition of 5 wt% irregularly-shaped 67%Ti-33%Fe particles which were less than 200 µm across [Chang 2010]. However, the choice of PtNP or AgNP additions strongly effected the observed change in the hardness and fracture toughness of porcelain. The H and KC were 4.94 ± 0.33 GPa and 1.36 ± 0.03 MPa·m½, respectively, for the porcelain matrix itself. The addition of AgNP raised the H and KC to 6.10 ± 0.14 GPa and 1.54 ± 0.05 MPa·m½, 23% and 13% greater, respectively, than the H and KC of the matrix. In contrast, after the addition of 26 wt% PtNP the hardness and fracture toughness were essentially unchanged from that of the dental porcelain matrix with H = 5.05 ± 0.15 GPa and KC = 1.42 ± 0.02 MPa·m½. Thus, while the addition of PtNP left both the H and KC values essentially intact, the addition of an equivalent weight % of AgNP to the same porcelain matrix material resulted in large changes in both H and KC [Fujieda 2012]. For dilute concentrations of nanoparticles, changes in the elastic modulus are typically on the order of a few percent or less. 6.3.4.2 Hardness and fracture toughness as function of porosity 6.3.4.2.1 Hardness as function of porosity In this study, the hardness, H, of Ba0.3Co4Sb12 and Ba0.3Co4Sb12-AgNP decreased linearly with increasing porosity (Figure 6.12a). In general, hardness is a function of porosity for ceramics or brittle semiconductors [Fan 2013; Ramadass 1983; Mangalaraja 2009]. A least-squares fit of the H versus volume fraction porosity, P, data was done to the relationship 168 H = H0(1 – bHP) (6) where H0 = the value of hardness at P = 0 and bH is a unitless constant that describes the rate of decrease in H with increasing P. For the four HP specimens, a least-squares fit of the hardness-porosity data to equation (6) yielded H0 = 4.3 ± 0.2 GPa and bH = 3.6 ± 0.2 with the coefficient of determination R2 = 0.84 (Figure 6.12a). For the three reprocessed specimens, H0 = 6.0 ± 0.5 GPa and bH = 3.0 ± 0.4, with R2 = 0.67 (Figure 6.12a). The low R2 is likely due to the small data set (three reprocessed specimens) and scatter in the data. 6.3.4.2.2 Fracture toughness as function of porosity The fracture toughness of the specimens in this study is independent of porosity, averaging 1.0 ± 0.2 MPa-m½ over the entire range of porosity for which KC was measured in in this study, namely from P = 0.05 to 0.21 (Figure 6.12b, Table 6.1). Note that fracture toughness was not measured for specimens PM-HP-673 (P = 0.30) or PM-HP673-Ag (P = 0.27) since these specimens were too fragile to be polished. In the literature, it has been noted that KC can increase, decrease, or remain unchanged with increasing P [Rice 1995; Shao 2013]. In a comprehensive review of the effects of porosity on the mechanical properties of brittle materials [Rice 1995], Rice noted, especially for the intermediate porosity range from roughly P = 0.1 to P = 0.15, there are a number of examples of materials in the literature (including Al2O3, B4C, Si3N4) showing “that KIC values for some of these porous bodies were clearly at least as high or higher than bodies with little or no porosity”. Proposed mechanisms for an increasing or constant KC with increasing 169 porosity include crack-pore interactions, where pores can blunt or deflect a growing crack [Case 2012b; Case 2012a]. 6.3.4.3 Hardness and fracture toughness as a function of load Over the indentation load range of 1.96 N to 9.8 N that is included in this study, both H and KC were independent of the load for Ba-skutterudite (Figure 6.12). The H, while independent of load, varied between specimens, but the KC was essentially constant for all seven specimens examined. The average KC of the seven specimens in this study is 1.0 ± 0.2 MPa-m½, which is comparable to the KC range of 0.46 MPa-m½ to 1.7 MPam½ measured for similar skutterudites (Table 6.7) [Ravi 2008; Eilertsen 2013; Rogl 2011]. For KC, the fracture toughness should ideally be independent of load. In order to test this load independence, the c3/2 (where c = radial crack length) versus indentation load, F, behavior was examined for each of the Ba-skutterudite specimens included in this study. A plot of c3/2 versus F yielded a straight line plot with R2 > 0.98 in each case, as demonstrated by plots from specimens VM-PECS-673 and PM-HP-873-Ag (Figure 6.13). Thus, the radial crack length versus indentation load behavior is consistent with equation (3) for each specimen in this study, where equation (3) can be rewritten as c3/ 2   ( E / H ) 0.5 F (7) KC For a given specimen, we assume the ξ, E, H and KC are constant, thus equation (7) is a straight line. The high value of R2 for the least-squares fit to equation (7) supports the use of equation (3) to calculate KC. 170 (a) Specimen VM-PECS-673 2 R = 0.997 3/2 500 c 1000 3/2 ( m ) 1500 0 0 2 4 6 8 10 F (N) 1500 (b) Specimen PM-HP-873-Ag 2 1000 500 c 3/2 3/2 ( m ) R = 0.998 0 0 2 4 6 8 F (N) 10 Figure 6.13. For each of the specimens, a plot of c3/2 versus load, F, was used to determine the suitability of the fracture toughness by Vickers indentation. For 5 of 7 specimens, the coefficient of determination, R2, for a linear regression to equation (7) through the data points is 0.99 or greater, with an R2 of 0.98 for specimen VM-PECS773, and 0.84 for specimen PM-HP-773. Note specimen PM-HP-773 also has a low fracture toughness measurement at 2.94 N load. All specimens were measured at three loads, except VM-PECS-673 and VM-PECS-773 with a fourth measurement at a Vickers indentation load of 1.96 N 171 In contrast to KC, hardness is frequently observed to be a function of the applied load. From the literature, for both metals [Pharr 2010; Nix 1998] and brittle solids [Bull 1989; Sangwal 2009], the measured hardness can be a function of the applied indentation load. The indentation size effect (ISE) [Nix 1998], where H decreases as the applied load increases, has been ascribed to elastic recovery and working hardening during indentation [Bull 1989; Sangwal 2009] while an observed reverse indentation size effect (RISE) [Sangwal 2000], where H decreases with decreasing load has been attributed to stress relaxation during unloading [Sangwal 2000]. 6.3.4.4 Hardness and fracture toughness results as a function of grain size The hardness varied as a function of grain size, as observed by the fit to equation (6) in Figure 6.12a, while the fracture toughness is independent of grain size, averaging 1.0 ± 0.2 MPa-m½ (Table 6.7). For the bimodal grain size HP specimens (matrix grain sizes of 0.3 µm to 3 µm, with larger grains approximately 5 µm to 20 µm across), the H0 is 4.3 ± 0.2 GPa, while for the reprocessed specimens with average grain sizes of 0.17 µm, 0.16 µm and 0.14 µm, the H0 is 6.0 ± 0.5 GPa (Figure 6.12a). 6.3.4.4.1 Hardness results as a function of grain size The difference in H0 between planetary milled and HP specimens and the vibratory milled and PECS processed specimens likely is related to the grain size dependence of H. A Hall-Petch type relationship between H and grain size, GS, available in the literature is [Rice 2000] H  H 0  kGS1 / 2 (8) 172 where H0 is considered as the H value in the single crystal limit and k is called the Petch parameter. While Hall-Petch behavior is very common in metals [Lawn 1993], the nature of the H versus GS dependence is uncertain for many ceramics [Rice 2000; Armstrong 2011]. Hall-Petch behavior is common for relatively soft ceramics such as alkaline halides, but there is a great deal of scatter in the H versus GS data for most hard ceramics [Rice 2000; Armstrong 2011]. Furthermore, the H0 obtained from H versus GS data if not always a good predictor for the H of single crystals (as is typically assumed for equation (8)) [Rice 2000]. In addition, for some ceramics, such as ZrO2 and AlMg2O4, H is independent of grain size [Rice 2000]. Also, minima in H versus GS data are often observed for intermediate grain sizes (GS on the order of several µm) such that the Petch parameter k can take on both negative and positive values (with unequal magnitudes) for a given data set [Rice 2000; Armstrong 2011]. Nevertheless, a well-accepted alternative to equation (8) is not available in the literature and thus, in general, the H versus GS behavior for brittle materials is not clear from the standpoint of the available data or the available functional relationship (equation (8)). 6.3.4.4.2 Fracture toughness results as a function of grain size In this study, fracture toughness, KC, is essentially independent of grain size, GS, averaging 1.0 ± 0.2 MPa-m½ (Figure 6.12b). This result may be counterintuitive, since fracture strength is typically a function of grain size [Barsoum 2003]. However, fracture strength is based on the stress required for failure of a specimen and is typically dependent on the largest flaw, and the intrinsic flaw size typically scales with GS [Barsoum 2003]. In contrast, KC is based on the work 173 necessary to extend a crack [Barsoum 2003], which may or may not be related to grain size. Thus, grain size engineering alone will not likely improve fracture toughness either for Ba-skutterudite or other cubic thermoelectric materials. Although no studies are currently available in the literature for the grain size dependence of KC for thermoelectric materials, KC has been observed to be essentially independent of grain size for various cubic materials, tested using a variety of test methods, including Y2O3 (notched beam test [Monroe 1978] and double cantilever beam technique [Rhoades 1986]), MgO (Chevron notch method [Yasuda 1990]), NiZn ferrite (notched beam [Veldkamp 1979]), and MgAl2O4 (double cantilever bean and Vickers indentation [Rice 1996]). In addition, for noncubic materials with small grain size, KC can be independent of grain size as was the case in a study by Yao et al. who found that KC was independent of grain size for PECS-processed, fully dense polycrystalline alumina with grain sizes ranging from 0.3 µm to 3.3 µm for KC measured by the surface crack in flexure technique [Yao 2011]. In contrast to cubic materials, for noncubic materials with intermediate grain sizes (roughly on the order of 10 to 100 µm), a maximum in KC has been observed as a function of grain size [Wachtman 2009]. Lawn and co-workers ascribed the observed increase in KC with increasing grain size to grain bridging [Vekinis 1990; Swanson 1987; Foulk III 2007], with the principle mechanism responsible for grain bridging being the clamping forces generated by thermal expansion anisotropy [Bennison 1989]. After the maximum in KC, KCMAX, is reached, the subsequent decrease in KC with further increases in grain size has been linked to thermal expansion anisotropy-induced microcracking in 174 noncubic materials, where the grain size at KCMAX coincides roughly with the critical grain size for microcracking [Rice 1998]. Thus, while for noncubic materials, a grainsize dependent maximum in KC has been observed, the insensitivity of KC to grain size for the cubic Ba-skutterudite in this study is consistent with the literature. In addition to the grain size and porosity trends, the value of KC for the Ba-skutterudite measured in this study is comparable to KC values in the literature for other CoSb3 systems [Ravi 2008; Eilertsen 2013; Rogl 2011] (Table 6.7). 6.4 Summary and conclusions In this study, the elastic moduli (Figure 6.10), hardness (Figure 6.12a), and fracture toughness (Figure 6.12b) of Ba0.3Co4Sb12 were insensitive to the addition of 0.5 wt% AgNP. Both the elastic moduli and hardness were functions of porosity (Figures 6.10, 6.12a), which agrees with previous literature results for elasticity (Figure 6.11) [Zhang 2010] and hardness [Ramadass 1983; Mangalaraja 2009; Shao 2013]. Fracture toughness (Figure 6.12b) was not a function of porosity (Figure 6.12b), which may indicate an interaction between the pores and the growing cracks. The study by Zhou [Zhou 2012] demonstrated that Ag addition results in important changes in thermoelectric properties (increasing electrical conduction) for Baskutterudite. However, this study indicates that the addition of silver also affects the densification process which in turn can affect changes in mechanical properties. At a sintering temperature of 873 K, the specimen without AgNP had a P value of 0.16 while the specimen sintered with 0.5 wt% AgNP had P = 0.05. The increased densification with the addition of Ag was likely due to a liquid phase sintering from the pertitectic reaction that formed the Ag3Sb phase observed in the specimen. 175 Based on the results of this study, a similar peritectic reaction may explain the formation of Ag3Sb and an enhanced ZT when Ag was added to ZnSb [Xiong 2013]. Similar peritectic or eutectic reactions may take place in other antimonide thermoelectric systems when Ag or another conductive material is added to enhance electrical conductivity. More generally, this type of eutectic or peritectic reaction to enhance densification and improve electrical conductivity may operate in a number of thermoelectric systems in addition to antimonides. Although a set of ternary or quaternary phase diagrams may not be available for many thermoelectric systems, a search of binary phase diagrams (such as the Ag-Sb phase diagrams used in this study) may act as a guide to discover potential beneficial peritectic or eutectic reactions. ACKNOWLEDGEMENTS The authors acknowledge the financial support of the Department of Energy, Revolutionary Materials for Solid State Energy Conversion Center, an Energy Frontiers Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under award number DE-SC0001054. 176 REFERENCES 177 REFERENCES [Alleno, Chen, Chubilleau, Lenoir, Rouleau, Trichet, Villeroy 2009] Alleno, E., Chen, L., Chubilleau, C., Lenoir, B., Rouleau, O., Trichet, M.F., Villeroy, B. Thermal Conductivity Reduction in CoSb3–CeO2 Nanocomposites. Journal of Electronic Materials 39 (2009) 1966–1970. [Androulakis, Lin, Kong, Uher, Wu, Hogan, Cook, Caillat, Paraskevopoulos, Kanatzidis 2007] Androulakis, J., Lin, C.-H., Kong, H.-J., Uher, C., Wu, C.-I., Hogan, T., Cook, B.A., Caillat, T., Paraskevopoulos, K.M., Kanatzidis, M.G. Spinodal decomposition and nucleation and growth as a means to bulk nanostructured thermoelectrics: enhanced performance in Pb(1-x)Sn(x)Te-PbS. Journal of the American Chemical Society 129 (2007) 9780–8. [Armstrong 2011] Armstrong, R.W. The Hardness and Strength Properties of WC-Co Composites. Materials 4 (2011) 1287–1308. [Barsoum 2003] Barsoum, M.W. Fundamentals of Ceramics. Taylor & Francis Group, New York, 2003. [Bedolla, Lemus-Ruiz, Contreras 2012] Bedolla, E., Lemus-Ruiz, J., Contreras, A. Synthesis and characterization of Mg-AZ91/AlN composites. Materials and Design 38 (2012) 91–98. [Bennison, Lawn 1989] Bennison, S.J., Lawn, B.R. Role of interfacial grain-bridging sliding friction in the crack-resistance and strength properties of nontransforming ceramics. Acta Metallurgica 37 (1989) 2659–2671. [Boccaccini 1994] Boccaccini, A. Comment on “Effective elastic moduli of porous ceramic materials.”Journal of the American Ceramic Society(1994) 2779–2781. [Bukat, Koscielski, Sitek, Jakubowska, Mlozniak 2011] Bukat, K., Koscielski, M., Sitek, J., Jakubowska, M., Mlozniak, A. Silver nanoparticles effect on the wettability of Sn-Ag-Cu solder pastes and solder joints microstructure on copper. Soldering & Surface Mount Technology 23 (2011) 150–160. [Bull, Page, Yoffe 1989] Bull, S.J., Page, T.F., Yoffe, E.H. An explanation of the indentation size effect in ceramics. Philosophical Magazine Letters 59 (1989) 281– 288. [Case 2012a] Case, E.D. Thermo-mechanical properties of thermoelectric materials. In: D.M. Rowe, ed., Thermoelectrics and its Energy Harvesting: Modules, Systems and Applications. CRC Press, Boca Raton, Florida, 2012, 581. 178 [Case 2012b] Case, E.D. Thermal Fatigue and Waste Heat Recovery via Thermoelectrics. Journal of Electronic Materials 41 (2012) 1811–1819. [Case, Smyth, Monthei 1981] Case, E.D., Smyth, J.R., Monthei, V. Grain-Size Determinations. Journal of the American Ceramic Society 64 (1981) C–24–C–25. [Chang, Chen, Ru, Yue, Yu, Zhang 2010] Chang, Q., Chen, D.L., Ru, H.Q., Yue, X.Y., Yu, L., Zhang, C.P. Toughening mechanisms in iron-containing hydroxyapatite/titanium composites. Biomaterials 31 (2010) 1493–501. [Cipriani, Corazza, Mazzetti 1996] Cipriani, C., Corazza, M., Mazzetti, G. Reinvestigation of natural silver antimonides. European Journal of Mineralogy 8 (1996) 1347–1350. [Couturier, Ducret, Merle, Disson, Joubert 1997] Couturier, R., Ducret, D., Merle, P., Disson, J.P., Joubert, P. Elaboration and characterization of a metal matrix composite: Al/AlN. Journal of the European Ceramic Society 17 (1997) 1861–1866. [Dharma, Shukor, Ariga 2009] Dharma, I.G.B.B., Shukor, M.H.A., Ariga, T. Wettability of Low Silver Content Lead-Free Solder Alloy. Materials Transactions 50 (2009) 1135–1138. [Dunn, Ledbetter 2011] Dunn, M.L., Ledbetter, H. Poisson’s ratio of porous and microcracked solids: Theory and application to oxide superconductors. Journal of Materials Research 10 (2011) 2715–2722. [Eilertsen, Subramanian, Kruzic 2013] Eilertsen, J., Subramanian, M. a., Kruzic, J.J. Fracture toughness of Co4Sb12 and In0.1Co4Sb12 thermoelectric skutterudites evaluated by three methods. Journal of Alloys and Compounds 552 (2013) 492–498. [Fan, Case, Yang, Nicholas 2013] Fan, X., Case, E.D., Yang, Q., Nicholas, J.D. Room temperature hardness of gadolinia-doped ceria as a function of porosity. Journal of Materials Science 48 (2013) 6977–6985. [Feschotte, Monachon, Durussel 1992] Feschotte, P., Monachon, F., Durussel, P. The binary system Sb-Ag: a revision of the Ag3Sb phase boundaries. Journal of Alloys and Compounds 186 (1992) L17–L18. [Foulk III, Cannon, Johnson, Klein, Ritchie 2007] Foulk III, J.W., Cannon, R.M., Johnson, G.C., Klein, P. a., Ritchie, R.O. A micromechanical basis for partitioning the evolution of grain bridging in brittle materials. Journal of the Mechanics and Physics of Solids 55 (2007) 719–743. [Fujieda, Uno, Ishigami, Kurachi, Wakamatsu, Doi 2012] Fujieda, T., Uno, M., Ishigami, H., Kurachi, M., Wakamatsu, N., Doi, Y. Addition of platinum and silver 179 nanoparticles to toughen dental porcelain. Dental Materials Journal 31 (2012) 711– 716. [Halpin 1992] Halpin, J.C. Primer on composite materials analysis. Technomic Publishing Company, Inc., Lancaster, Pennsylvania, 1992. [Hashin 1962] Hashin, Z. The Elastic Moduli of Heterogeneous Materials. Journal of Applied Mechanics 29 (1962) 143–150. [Hassam, Bahari, Legendre 2001] Hassam, S., Bahari, Z., Legendre, B. Phase diagram of the Ag–Bi–Sb ternary system. Journal of Alloys and Compounds 315 (2001) 211– 217. [He, Hu, Tang, Lan, Yang, Wang, Ren, Hao, Chen 2008] He, Q., Hu, S., Tang, X., Lan, Y., Yang, J., Wang, X., Ren, Z., Hao, Q., Chen, G. The great improvement effect of pores on ZT in Co[sub 1−x]Ni[sub x]Sb[sub 3] system. Applied Physics Letters 93 (2008) 042108. [Hutton 2004] Hutton, D. V. Fundamentals of finite element analysis. McGraw-Hill, New York, 2004. [Ioffe 1960] Ioffe, A. Physics of Semicondutors. Infosearch Limited, London, 1960. [Ji, He, Alboni, Su, Gothard, Zhang, Tritt, Kolis 2007] Ji, X., He, J., Alboni, P., Su, Z., Gothard, N., Zhang, B., Tritt, T.M., Kolis, J.W. Thermal conductivity of CoSb3 nano-composites grown via a novel solvothermal nano-plating technique. physica status solidi RRL – Rapid Research Letters 1 (2007) 229–231. [Kim, Bush 1999] Kim, H.S., Bush, M.B. The effects of grain size and porosity on the elastic modulus of nanocrystalline materials. Nanostructured Materials 11 (1999) 361–367. [Lawn 1993] Lawn, B.R. Fracture of Brittle Solids. Cambridge University Press, New York, 1993. [Lawn, Cook 2012] Lawn, B.R., Cook, R.F. Probing material properties with sharp indenters: a retrospective. Journal of Materials Science 47 (2012) 1–22. [Lee, Vashaee, Wang, Dresselhaus, Ren, Chen 2010] Lee, H., Vashaee, D., Wang, D.Z., Dresselhaus, M.S., Ren, Z.F., Chen, G. Effects of nanoscale porosity on thermoelectric properties of SiGe. Journal of Applied Physics 107 (2010) 094308. [Mangalaraja, Ananthakumar, Uma, Jiménez, López, Camurri 2009] Mangalaraja, R.V., Ananthakumar, S., Uma, K., Jiménez, R.M., López, M., Camurri, C.P. 180 Microhardness and fracture toughness of Ce0.9Gd0.1O1.95 for manufacturing solid oxide electrolytes. Materials Science and Engineering: A 517 (2009) 91–96. [Mi, Zhao, Zhu, Tu 2008] Mi, J.L., Zhao, X.B., Zhu, T.J., Tu, J.P. Thermoelectric properties of Yb 0.15 Co 4 Sb 12 based nanocomposites with CoSb 3 nanoinclusion. Journal of Physics D: Applied Physics 41 (2008) 205403. [Migliori, Sarrao 1997] Migliori, A., Sarrao, J.L. Resonant Ultrasound Spectroscopy: Applications to Physics, Materials Measurements, and Nondestructive Evaluation. Wiley-Vch, New York, 1997. [Monroe, Smyth 1978] Monroe, L.D., Smyth, J.R. Grain Size Dependence of Fracture Energy of Y2O3. Journal of the American Ceramic Society 61 (1978) 538–539. [Morelli, Meisner 1995] Morelli, D.T., Meisner, G.P. Low temperature properties of the filled skutterudite CeFe4Sb12. Journal of Applied Physics 77 (1995) 3777. [Ni, Ren, Case, Timm 2009] Ni, J., Ren, F., Case, E., Timm, E. Porosity dependence of elastic moduli in LAST (lead–antimony–silver–tellurium) thermoelectric materials. Materials Chemistry and Physics(2009) . [Ni, Case, Khabir, Stewart, Wu, Hogan, Timm, Girard, Kanatzidis 2010] Ni, J.E., Case, E.D., Khabir, K.N., Stewart, R.C., Wu, C.-I., Hogan, T.P., Timm, E.J., Girard, S.N., Kanatzidis, M.G. Room temperature Young’s modulus, shear modulus, Poisson's ratio and hardness of PbTe–PbS thermoelectric materials. Materials Science and Engineering: B 170 (2010) 58–66. [Nix, Gao 1998] Nix, W.D., Gao, H. Indentation size effects in crystalline materials: A law for strain gradient plasticity. Journal of the Mechanics and Physics of Solids 46 (1998) 411–425. [Nolas, Kaeser, Littleton, Tritt 2000] Nolas, G.S., Kaeser, M., Littleton, R.T., Tritt, T.M. High figure of merit in partially filled ytterbium skutterudite materials. Applied Physics Letters 77 (2000) 1855. [Nolas, Morelli, Tritt 1999] Nolas, G.S., Morelli, D.T., Tritt, T.M. SKUTTERUDITES: A Phonon-Glass-Electron Crystal Approach to Advanced Thermoelectric Energy Conversion Applications. Annual Review of Materials Science 29 (1999) 89–116. [Okamoto 1993] Okamoto, H. Ag-Sb (Silver-Antimony). Journal of Phase Equilibria 14 (1993) 531–532. [Okamoto 2007] Okamoto, H. Ag-Sb (Silver-Antimony). Journal of Phase Equilibria and Diffusion 28 (2007) 403–403. 181 [Pharr, Herbert, Gao 2010] Pharr, G.M., Herbert, E.G., Gao, Y. The Indentation Size Effect: A Critical Examination of Experimental Observations and Mechanistic Interpretations. Annual Review of Materials Research 40 (2010) 271–292. [Pilchak, Ren, Case, Timm, Schock, Wu, Hogan 2007] Pilchak, A.L., Ren, F., Case, E.D., Timm, E.J., Schock, H.J., Wu, C.-I., Hogan, T.P. Characterization of dry milled powders of LAST (lead–antimony–silver–tellurium) thermoelectric material. Philosophical Magazine 87 (2007) 4567–4591. [Ramadass, Mohan, Reddy 1983] Ramadass, N., Mohan, S., Reddy, S.R. Studies on the metastable phase retention and hardness in zirconia ceramics. Materials Science and Engineering 60 (1983) 65–72. [Ramakrishnan, Arunachalamt 1993] Ramakrishnan, N., Arunachalamt, V.S. Effective Elastic Moduli of Porous Ceramic Materials. Journal of the American Ceramic Society 76 (1993) 2745–2752. [Ravi, Firdosy, Caillat, Brandon, Walde, Maricic, Sayir 2009] Ravi, V., Firdosy, S., Caillat, T., Brandon, E., Walde, K., Maricic, L., Sayir, A. Thermal Expansion Studies of Selected High-Temperature Thermoelectric Materials. Journal of Electronic Materials 38 (2009) 1433–1442. [Ravi, Firdosy, Caillat, Lerch, Calamino, Pawlik, Nathal, Sechrist, Buchhalte, Nutt 2008] Ravi, V., Firdosy, S., Caillat, T., Lerch, B., Calamino, A., Pawlik, R., Nathal, M., Sechrist, A., Buchhalte, J., Nutt, S. Mechanical Properties of Thermoelectric Skutterudites. Space Technology and Applications International Forum—STAIF 2008(2008) 656–662. [Ren, Case, Ni, Timm, Lara-Curzio, Trejo, Lin, Kanatzidis 2009] Ren, F., Case, E.D., Ni, J.E., Timm, E.J., Lara-Curzio, E., Trejo, R.M., Lin, C.-H., Kanatzidis, M.G. Temperature-dependent elastic moduli of lead telluride-based thermoelectric materials. Philosophical Magazine 89 (2009) 143–167. [Ren, Case, Timm, Schock 2008] Ren, F., Case, E.D., Timm, E.J., Schock, H.J. Hardness as a function of composition for n-type LAST thermoelectric material. Journal of Alloys and Compounds 455 (2008) 340–345. [Rhoades, Baldoni, Wei 1986] Rhoades, W., Baldoni, J., Wei, G. Final report for ORN contract N00014-82-C-0452. GTE Laboratory(1986) . [Rice 1995] Rice, R.W. Comment on “Effective elastic moduli of porous ceramic materials.”Journal of the American Ceramic Society 78 (1995) 1711. [Rice 1996] Rice, R.W. Grain size and porosity dependence of ceramic fracture energy and toughness at 22°C. Journal of Materials Science 31 (1996) 1969–1983. 182 [Rice 1998] Rice, R.W. Porosity of Ceramics. Marcel Dekker, New York, 1998. [Rice 2000] Rice, R.W. Mechanical Properties of Ceramics and Composites: Grain And Particle Effects. Marcel Dekker, New York, 2000. [Rogl, Rogl 2011] Rogl, G., Rogl, P. Mechanical Properties of Skutterudites. Science of Advanced Materials 3 (2011) 517–538. [Sangwal 2000] Sangwal, K. On the reverse indentation size effect and microhardness measurement of solids. Materials Chemistry and Physics 63 (2000) 145–152. [Sangwal 2009] Sangwal, K. Review: Indentation size effect, indentation cracks and microhardness measurement of brittle crystalline solids - some basic concepts and trends. Crystal Research and Technology 44 (2009) 1019–1037. [Schmidt, Case, Lehr, Morelli 2013] Schmidt, R.D., Case, E.D., Lehr, G.J., Morelli, D.T. Room temperature mechanical properties of polycrystalline YbAl3, a promising low temperature thermoelectric material. Intermetallics 35 (2013) 15–24. [Schmidt, Ni, Case, Sakamoto, Kleinow, Wing, Stewart, Timm 2010] Schmidt, R.D., Ni, J.E., Case, E.D., Sakamoto, J.S., Kleinow, D.C., Wing, B.L., Stewart, R.C., Timm, E.J. Room temperature Young’s modulus, shear modulus, and Poisson's ratio of Ce0.9Fe3.5Co0.5Sb12 and Co0.95Pd0.05Te0.05Sb3 skutterudite materials. Journal of Alloys and Compounds 504 (2010) 303–309. [Segerland 1984] Segerland, L. Applied finite element analysis. Wiley, New York, 1984. [Shao, Du, Wu, Song, Xu, Jiang 2013] Shao, Y., Du, R., Wu, X., Song, F., Xu, X., Jiang, C. Effect of porosity on the crack pattern and residual strength of ceramics after quenching. Journal of Materials Science 48 (2013) 6431–6436. [Shi, Yang, Salvador, Chi, Cho, Wang, Bai, Yang, Zhang, Chen 2011] Shi, X., Yang, J., Salvador, J.R., Chi, M., Cho, J.Y., Wang, H., Bai, S., Yang, J., Zhang, W., Chen, L. Multiple-filled skutterudites: high thermoelectric figure of merit through separately optimizing electrical and thermal transports. Journal of the American Chemical Society 133 (2011) 7837–46. [Slack, Tsoukala 1994] Slack, G. a., Tsoukala, V.G. Some properties of semiconducting IrSb3. Journal of Applied Physics 76 (1994) 1665–1671. [Swanson, Fairbanks, Lawn, Mai, Hockey 1987] Swanson, P.L., Fairbanks, C.J., Lawn, B.R., Mai, Y.-W., Hockey, B.J. Crack-Interface Grain Bridging as a Fracture Resistance I, Mechanism in Ceramics: I, Experimental Study on Alumina. Journal of the American Ceramic Society 70 (1987) 279–289. 183 [Tokiai, Uesugi, Nosaka, Hirayama, Ito, Koumoto 1997] Tokiai, T., Uesugi, T., Nosaka, M., Hirayama, A., Ito, K., Koumoto, K. Thermoelectric properties of Mn-doped iron disilicide ceramics fabricated from radio-frequency plasma-treated fine powders. Journal of materials Science 32 (1997) 3007–3011. [Toprak, Stiewe, Platzek, Williams, Bertini, Muller, Gatti, Zhang, Rowe, Muhammed 2004] Toprak, M.S., Stiewe, C., Platzek, D., Williams, S., Bertini, L., Muller, E., Gatti, C., Zhang, Y., Rowe, M., Muhammed, M. The Impact of Nanostructuring on the Thermal Conductivity of Thermoelectric CoSb3. Advanced Functional Materials 14 (2004) 1189–1196. [Uher 2001] Uher, C. Skutterudites: Prospective Novel Thermoelectrics. In: T.M. Tritt, ed., Semiconductors and Semimetals, Vol. 69. Academic Press, San Diego, 2001, 139–253. [Underwood 1969] Underwood, E.E. Stereology, or the quantitative evaluation of microstructures. Journal of Microscopy 89 (1969) 161–180. [Vekinis, Ashby, Beaumont 1990] Vekinis, G., Ashby, M.F., Beaumont, P.W.R. R-curve behaviour of Al2O3 ceramics. Acta Metallurgica et Materialia 38 (1990) 1151– 1162. [Veldkamp, Hattu 1979] Veldkamp, J., Hattu, N. On the fracture toughness of brittle materials. Philips Journal Research 34 (1979) 1–25. [Voronin, Osadchii 2013] Voronin, M. V., Osadchii, E.G. Standard thermodynamic properties of Ag3Sb and Ag6Sb evaluated by EMF measurements. Inorganic Materials 49 (2013) 550–554. [Wachtman, Cannon, Matthewson 2009] Wachtman, J.B., Cannon, W.R., Matthewson, M.J. Mechanical Properties of Ceramics. Wiley-Vch, Hoboken, N.J., 2009. [Wen, Duan, Zhai, Li, Zhang 2013] Wen, P., Duan, B., Zhai, P., Li, P., Zhang, Q. Effect of thermal annealing on the microstructure and thermoelectric properties of nanoTiN/Co4Sb11.5Te0.5 composites. Journal of Materials Science: Materials in Electronics 24 (2013) 5155–5161. [Xiong, Okamoto, Inui 2013] Xiong, D.-B., Okamoto, N.L., Inui, H. Enhanced thermoelectric figure of merit in p-type Ag-doped ZnSb nanostructured with Ag3Sb. Scripta Materialia 69 (2013) 397–400. [Yang, Wu, Zhang 2004] Yang, L., Wu, J.S., Zhang, L.T. Synthesis of filled skutterudite compound La0.75Fe3CoSb12 by spark plasma sintering and effect of porosity on thermoelectric properties. Journal of Alloys and Compounds 364 (2004) 83–88. 184 [Yao, Liu, Holland, Huang, Xiong, Schoenung, Mukherjee 2011] Yao, W., Liu, J., Holland, T.B., Huang, L., Xiong, Y., Schoenung, J.M., Mukherjee, A.K. Grain size dependence of fracture toughness for fine grained alumina. Scripta Materialia 65 (2011) 143–146. [Yasuda, SD, Kanemichi, Matsuo, Kimura 1990] Yasuda, K., SD, K., Kanemichi, Y., Matsuo, Y., Kimura, S. Influence of Grain Size and Temperature on Fracture Toughness of MgO Sintered Body. Journal of the Ceramic Society of Japan 98 (1990) 1103–1108. [Yoon, Kwon, Ahn, Kim, Koo, Bae, Cho, Kim, Park 2013] Yoon, S., Kwon, O.-J., Ahn, S., Kim, J.-Y., Koo, H., Bae, S.-H., Cho, J.-Y., Kim, J.-S., Park, C. The Effect of Grain Size and Density on the Thermoelectric Properties of Bi2Te3-PbTe Compounds. Journal of Electronic Materials 42 (2013) 3390–3396. [Yoshizawa, Nakanishi, Kumagai, Oikawa, Sekine, Shirotani 2004] Yoshizawa, M., Nakanishi, Y., Kumagai, T., Oikawa, M., Sekine, C., Shirotani, I. Elastic Anomalies of Polycrystalline SmRu 4 P 12 Associated with Metal–Insulator Transition. Journal of the Physical Society of Japan 73 (2004) 315–318. [Zawrah, Abdel-kader, Elbaly 2012] Zawrah, M.F., Abdel-kader, H., Elbaly, N.E. Fabrication of Al2O3–20vol.% Al nanocomposite powders using high energy milling and their sinterability. Materials Research Bulletin 47 (2012) 655–661. [Zhang, Rogl, Grytsiv, Puchegger, Koppensteiner, Spieckermann, Kabelka, Reinecker, Rogl, Schranz, Zehetbauer, Carpenter 2010] Zhang, L., Rogl, G., Grytsiv, A., Puchegger, S., Koppensteiner, J., Spieckermann, F., Kabelka, H., Reinecker, M., Rogl, P., Schranz, W., Zehetbauer, M., Carpenter, M.A. Mechanical properties of filled antimonide skutterudites. Materials Science and Engineering: B 170 (2010) 26–31. [Zhang, Sakamoto 2013] Zhang, L., Sakamoto, J. The microstructural stability and thermoelectric properties of Mm0.9Fe3.5Co0.5Sb12-based skutterudites. Materials Chemistry and Physics 138 (2013) 601–607. [Zhao, He, Hao, Wu, Hogan, Wolverton, Dravid, Kanatzidis 2012] Zhao, L.-D., He, J., Hao, S., Wu, C.-I., Hogan, T.P., Wolverton, C., Dravid, V.P., Kanatzidis, M.G. Raising the thermoelectric performance of p-type PbS with endotaxial nanostructuring and valence-band offset engineering using CdS and ZnS. Journal of the American Chemical Society 134 (2012) 16327–36. [Zhou, Li, Kita 2008] Zhou, M., Li, J.-F., Kita, T. Nanostructured AgPb(m)SbTe(m+2) system bulk materials with enhanced thermoelectric performance. Journal of the American Chemical Society 130 (2008) 4527–32. 185 [Zhou, Wang, Zhang, Chi, Su, Sakamoto, Uher 2012] Zhou, X., Wang, G., Zhang, L., Chi, H., Su, X., Sakamoto, J., Uher, C. Enhanced thermoelectric properties of Bafilled skutterudites by grain size reduction and Ag nanoparticle inclusion. Journal of Materials Chemistry 22 (2012) 2958. [Zienkiewicz, Taylor 2005] Zienkiewicz, O.C., Taylor, R.L. The Finite Element Method for Solid and Structural Mechanics. Butterworth-Heinemann, Boston, 2005. 186 7 High temperature elastic moduli of thermoelectric SnTe1+X –y SiC nanoparticulate composites Robert D. Schmidt 1, Eldon D. Case 1, Jennifer E. Ni1, Rosa M. Trejo 2, Edgar LaraCurzio2, Rachel J. Korkosz3, Mercouri G. Kanatzidis3 1 Chemical Engineering and Materials Science Department, Michigan State University, East Lansing, MI, 48824 2 High Temperature Materials Laboratory, Oak Ridge National Laboratory, Oak Ridge, TN 37831 3 Department of Chemistry, Northwestern University, Evanston, IL, 60208 Published in: Journal of Materials Science 48 (2013) 8244 – 8258. Abstract In waste heat recovery applications, thermoelectric (TE) generators are subjected to thermal gradients and thermal transients, creating mechanical stresses in the thermoelectric legs. Such stresses are functions of the elastic moduli of the TE material. For SnTe1+X matrices (where x = 0.0 or 0.016) composite specimens with 0 to 4 vol.% SiC nanoparticle (SiCNP) additions, the elastic moduli (Young’s modulus, shear modulus, bulk modulus and Poisson’s ratio) were measured by resonant ultrasound spectroscopy (RUS) from room temperature to 663 K. The effects of matrix composition and the SiCNP additions on the room temperature intercepts and the slopes of the elastic modulus as a function of temperature are also discussed. Keywords: thermoelectrics; elastic modulus; SnTe; SiC nanoparticles 187 7.1 Introduction A number of thermoelectric (TE) materials, including SnTe are candidates for harvesting waste heat from a variety of sources, including incinerators and automotive or truck exhaust. The energy conversion efficiency, η, of a thermoelectric device can be written as [Bux 2010]    TH  TC   1  ZT  1       TH   1  ZT   TC T   H          (1) where TH and TC are, respectively, the hot and cold side temperatures of the thermoelectric device and T is the average device temperature. ZT is the dimensionless figure of merit which is given by ZT  S 2 eT (2)  where S = the Seebeck coefficient, σe = the electrical conductivity, S2 σe = the power factor and κ = the thermal conductivity. As shown by equation (1), as ZT increases the conversion efficiency increases. However, in addition to the thermal and electrical transport properties, the mechanical properties of the thermoelectric material are also important. Especially in the context of harvesting waste energy, as the heat source (say an automobile engine or an industrial furnace) is cycled on and off, the TE elements that are in place to capture waste heat will also experience thermal transients. The maximum surface thermal stress, max, induced by a thermal transient on a flat plate of uniform thickness is a function of the 188 temperature-dependent elastic modulus, E(T), and Poisson’s ratio, (T), and thermal expansion coefficient, (T) such that [Lu 1998; Zhao 2000; Manson 1966] max  E T  αT   T  f(Bi) 1   T  (3) The function f(Bi) increases monotonically with increasing Biot modulus Bi, where Bi = ah/κ [Kreith 1973] and a is the characteristic dimension of the specimen, h is the surface heat transfer coefficient and κ is the thermal conductivity of the specimen . In the thermal transient case, T is the difference between the initial and final temperature during the temperature excursion and the function f(Bi) depends on the dimensionless parameter Bi. Also, T = Ti  T  is the quench temperature difference, where Ti is the initial temperature of the specimen and T∞ is the temperature of the ambient medium. In addition to stresses induced by thermal transients (equation 3), there are thermal gradients imposed between the hot and cold plates of a thermoelectric device that are essential for current to flow [D’Angelo 2011]. Also, the thermoelectric elements will be subjected to stresses due to thermal expansion mismatch within the thermoelectric module as well as to various external, mechanical stresses. Regardless of the mechanism that generates the stress, calculations via both finite element [Martin 1973; Kaliakin 2002] and analytical techniques [Manson 1966] require knowledge of the elastic modulus and Poisson’s ratio in order to analyze the stress-strain response of the material. Also, if microcracks are induced by either mechanical or thermal loading, then the microcrackinduced changes in elastic moduli provide a means of monitoring the extent of microcrack damage [Fan 2012; Case 1993; Chotard 2008]. 189 In this study, the Young’s modulus, shear modulus, Poisson’s ratio, acoustic wavespeed and Debye temperature were measured over the temperature range from 298 K to 663 K for six bulk SnTe1+X -SiC nanoparticle composites specimens with the SiC nanoparticle (SiCNP) loading from 0 to 4 volume percent. Also, the effects of an offstoichiometric SnTe composition on the elastic moduli are discussed. 7.2 Experimental procedure 7.2.1 Specimen Preparation To fabricate each SnTe1+X ingot, high purity powders of Sn (Plasmaterials, purity 99.999%) and Te (5N Plus Inc., purity 99.999%) with no added dopants were flame sealed in a quartz tube under vacuum. For ingot SnTe(x=0) , stoichiometric amounts of each powder were placed into quartz tubes (21.2499 g of Sn and 22.8462 g of Te in first batch, 21.2478 g of Sn and 22.8486 g of Te in second batch). The sealed tubes were heated to 1273 K in 10 hours, held at 1273 K for 10 hours then air quenched to room temperature (RT). For ingot SnTe(x=0.016), a non-stoichiometric amount of powder was added (44.9274 g of Sn and 49.0715 g of Te) to a quartz tube to produce SnTe1.016. The sealed tube was heated to 1148 K in 12 hours, held for 12-13 hours at temperature, then quenched into water. The resulting ingot was crushed and ground in a WC-lined mechanical mortar and pestle. The ground powder was passed through a 53 µm sieve on a shaker table for 5 to 15 minutes. Powder that did not pass through the sieve was reground and re-sieved. Additional details on the crush-grind-sieve-regrind (CGSR) procedure are provided elsewhere [Pilchak 2007]. 190 Table 7.1. For each of the six SnTe-SiCNP specimens included in this study, the volume fraction SiCNP addition, specimen mass, dimensions, mass density, porosity, P, and mean grain size. Specimens were prepared from one of two starting ingots, designated A or B in the specimen label. Mass (g) Dimensions (cm) Mass P Density (g/cm3) SnTe(x=0)V0 SnTe(x=0)V0.01 SnTe(x=0)V0.02 SiC Nanoparticle Volume Fraction 0.00 0.01 0.02 7.7749 7.4952 7.3529 6.27 6.09 6.21 SnTe(x=0)V0.03 0.03 3.4264 6.13 0.034 1.0 SnTe(x=0)V0.04 0.04 SnTe(x=0.016)V0 0.00 a Not available 7.6725 7.5693 Ø 1.984 × 0.401 Ø 1.973 × 0.403 Ø 1.989 × 0.31 1.162 × 1.249 × 0.385 Ø 1.968 × 0.416 Ø 1.987 × 0.387 Mean Grain Size (µm) 0.028 N.A. a 0.051 2.7 0.026 1.0 6.07 6.30 0.039 1.2 0.022 1.8 Temperature Pressure Temperature (K) 700 60 600 40 500 400 20 300 0 100 200 300 Pressure (MPa) Specimen label 0 Time (min) Figure 7.1. Processing parameters (temperature and die pressure) as a function of time for the PECS densification of the SnTe-SiCNP specimens included in this study. 191 After grinding, the CGSR powder was milled in a planetary ball mill in 20 g to 30 g batches with 100 g of 10 mm diameter spherical alumina media at 150 RPM for 3 h (PM100, Retsch GmbH, Haan, Germany). For specimens with SiC nanoparticle (SiCNP) additions, 95% purity silicon SiCNP with a vendor –specified particle size of 50-60 nm (SiCNP, 4621HW, Lot #4621-110209, Nanostructured & Amorphous Materials Inc., Houston, TX) was added to an 8 g batch of milled SnTe1+X powder and mixed at 110 RPM for 6 h with 100 g of 10 diameter mm alumina media. From 0 to 4 vol% SiCNP was added to the SnTe1+X powders (Table 7.1). Sintering of the milled powder was performed by pulsed electric current sintering (PECS) in a 20 mm diameter graphite die lined with two layers of grafoil (SPS Model 103, Thermal Technology LLC, Santa Rosa, CA). For each specimen included in this study, the PECS sintering was performed using an argon-filled chamber at 673 K and at a die pressure of 60 MPa for 20 minutes, followed by 210 minutes at 573 K. Details of the time-temperature and pressure-temperature profiles are given in Figure 7.1. The resulting disk-shaped specimens were from about 3 to 4 mm thick, although one specimen (A-043) was cut into a rectangular parallelepiped prior to testing (Table 7.1). 7.2.2 Specimen characterization The microstructure of the specimens was examined using the following scanning electron microscopes (JEOL 6400, JEOL 6610LV or JSM-7500F, JEOL Ltd., Japan). All SEM images taken with the JEOL 6400, JEOL 6610LV used 15 mm working distance with a 15 kV accelerating voltage. For the JSM-7500F (high resolution SEM) a 4.5 mm working distance and 2 kV accelerating voltage was used when imaging powder, 4.5 mm working distance and 5 kV when imaging bulk specimens by secondary electron only, 192 and 8 mm working distance and 15 kV accelerating voltage on specimens imaged by backscatter electron and for energy-dispersive X-ray spectroscopy (EDS). The JSM-7500F was used to image fractured and polished surfaces of the SnTe1+X bulk specimens along with the as-received and as-ground SiCNP. The fractured surfaces of the specimens were not coated prior to SEM observation. However, the polished surfaces and the SiCNP powders themselves were sputter coated with osmium since without coating the specimens exhibited localized charging. In order to attempt to break up extensive clumping in the as-received SiCNP, a 5 gram batch of the as-received SiCNP powders was mixed with 5 ml of ethanol and then hand ground for 5 minutes with an alumina mortar and pestle. While the grinding process in the mortar and pestle is certainly not identical to the grinding process experienced during planetary milling of the SiCNP and SnTe1+X powder mixture that was subsequently sintered, the hand grinding gives a rough gauge of the how easy it is to break-up the agglomerates. The milled powder was examined by XRD (Miniflex II, Cu Kα radiation, Rigaku, Tokyo, Japan) to verify the phase composition and lattice parameter of the SnTe1+X material. 7.2.3 High temperature Resonant Ultrasound Spectroscopy measurements The elastic moduli were measured by Resonant Ultrasound Spectroscopy (RUS) in a furnace with a flowing Ar-4% H atmosphere. The elastic moduli were measured at RT before turning on the furnace. The specimen was heated to 333 K for the first measurement at higher temperature, then the furnace temperature was increased in 30 K increments to 663 K. The specimen was then cooled in 30 K increments from 663 K to 193 333 K. After the furnace was turned off and the specimen cooled to RT, the specimen was measured one final time by RUS. Dimensional changes during heating were calculated using a coefficient of thermal expansion of 22.2 x 10-6 K-1 (Schmidt et al., unpublished data). Additional details of the HT-RUS procedures are provided elsewhere [Ren 2008; Ren 2009b; Ren 2009a]. 7.3 Results and Discussion 7.3.1 Microstructural analysis Both fracture surfaces (Figure 7.2a) and polished surfaces (Figure 7.2b) of the SnTe1+X -SiCNP composite specimens were examined. No surface-breaking cracks were observed for either the fractured surfaces (Figure 7.2a) or the polished surfaces (Figure 7.2b) of the specimens. The polished surfaces of the SnTe1+X specimens showed occasional isolated, quasi-spherical pores with submicron diameters (Figure 7.2b). In addition, the grains were relatively equiaxed with a mean grain size of roughly 1 – 3 µm (Figure 7.2a, Table 7.1), as evaluated by the linear intercept technique on SEM micrographs of fractured surfaces [Case 1981b]. Also, the fractures were predominantly intergranular with transgranular fracture observed for some of the larger grains (Figure 7.2a). The mass, mass density, grain size and volume fraction porosity for each specimen included in this study is listed in Table 7.1. Room temperature X-ray diffraction showed that only the SnTe cubic phase was present in the specimens from both ingots SnTe(x=0) and SnTe(x=0.016) (Table 7.1), however there was a systematic shift in the XRD patterns that was likely due to compositional differences in SnTe1+X solid solution. The lattice parameter was a0 = 6.3130 + 0.0056 Å for the planetary-milled powders used to fabricate specimens each of 194 Figure 7.2. For SnTe specimen A-06 (0 vol% SiCNP), (a) fractured specimen surface and (b) polished specimen surface. Figure 7.3. The as-received SiCNP consisted of particles roughly 50 nm across agglomerated into clusters ranging in size from approximately 100 nm to 20 µm. Figure 7.4. (a) After grinding in an alumina mortar and pestle for 5 min in ethanol, the SiCNP exhibited agglomerations of particles ranging from sub-micron to ~10 µm diameter. (b) The agglomerations consisted primarily of nanoparticles with diameter less than 100 nm, but with an occasional micron-scale particle. The size distribution of the agglomerates after manual grinding was similar to that observed prior to manual grinding (Figure 7.3). 195 the “SnTe(x=0.0)” specimens, of nominal composition SnTe. For the planetary-milled powders used to fabricate the “SnTe(x=0.016)” specimen, of composition SnTe1.016, the lattice parameter was 6.3014 + 0.0056 Angstroms. The literature value for lattice parameter of SnTe given by is 6.313 Angstroms [Wyckoff 1963]. As Te is added to the SnTe structure, Sn vacancies are created in the rock salt lattice, so that the shift in stoichiometry results from the creation of cation vacancies [Rogacheva 1991]. The compositional difference between ingots SnTe(x=0) and SnTe(x=0.016) has important implications for mechanical properties of the SnTe1+X specimens since in general the elastic moduli are known to be sensitive to changes in composition within solid solution systems [Ravinder 2001; Srinivas Rao 2003; Ren 2007] including thermoelectric solid solutions [Ren 2007]. The composition-related changes in elastic moduli are discussed further in following section. The as-received SiCNP powders were not discrete nanoparticles but instead consisted of clumps with diameters from roughly 100 nm to 20 µm across (Figure 7.3), with individual particles roughly 100 nm across (Figure 7.3). Manual grinding in a mortar and pestle did not noticeably reduce the clump size from the original several hundred nm diameter (Figure 7.4). The manual grinding only provides a rough gauge of the ease of breaking up the SiCNP powder agglomerates and does not approximate the grinding action of the planetary mill used to process the bulk specimens in this study. Fracture surfaces of specimens SnTe(x=0)V0, SnTe(x=0)V0.01, SnTe(x=0)V0.02, SnTe(x=0)V0.03 and SnTe(x=0)V0.04 were examined (Figure 7.5a – 7.5e). Each specimen with SiCNP addition showed fine-grained clusters on the fracture surface that likely represented SiCNP clumps on the surface that ranged from about 196 Figure 7.5. Fractured specimen surface images for (a) specimen SnTe (x = 0)V0 with no SiCNP added, and for (b) SnTe (x = 0)V0.01, (c) SnTe(x = 0)V0.02, (d) SnTe(x = 0)V0.03, and (e) SnTe(x = 0)V0.04. In (b) – (e), the location of SiCNP clusters is indicated by arrows. 197 200 nm to 2000 nm in across (Figure 7.5b – 7.5e). The SiCNP clumps are indicated by arrows in Figures 7.5b – 7.5e. The fracture surface of the SnTe(x=0)V0 was free of such clusters. In order to determine whether or not the clusters observed on the specimen fracture surfaces of the composites were composed of SiCNP, pairs of secondary electron and backscattered images were taken at high resolution (Figures 7.6a – 7.6d). For specimen SnTe(x=0)V0, with no SiCNP addition, the secondary electron image (Figure 7.6a) and the backscattered image (Figure 7.6b) were similar in appearance. However, for SnTe(x=0)V0.04 (with 0.04 volume fraction of SiCNP), the backscattered image of the specimen exhibited dark regions that indicate a lower atomic mass, which is consistent with the presence of SiCNP at or just below the surface of the specimen in the dark regions (Figure 7.6d). However, for the secondary electron image (Figure 7.6d), there was much less atomic number contrast in the image. From measurements on SEM micrographs taken at two different magnifications, the area fraction of SiCNP clusters was estimated as roughly 2.1–2.5 % (Table 7.2). EDS measurements on specimen SnTe(x = 0)V0.04 indicated a silicon content of about 13 at% within dark region (the box-labeled ‘‘Spectrum 2’’ in Figure 7.7). In contrast, a Si content of <1 at% was indicated by EDS in the light region marked by the box-labeled ‘‘Spectrum 1’’ in Figure 7.7. Thus, there is a very strong indication that SiCNP clusters ranging from roughly 100 to 1000 nm are distributed within specimen SnTe(x = 0)V0.04. 198 Figure 7.6. Specimen SnTe(x = 0)V0, with no added SiCNP, shows no surface features other than surface debris in both (a) a secondary electron image (SEI) and in (b) a backscattered electron image. Specimen SnTe(x = 0)V0.04 with 0.04 vol fraction SiCNP addition shows a difference between the (c) secondary electron image and (d) the backscattered electron image, with the image in the backscattered electron mode showing dark regions indicating likely SiCNP clusters at or near the specimen surface. 199 Table 7.2. For SiCNP clusters identified by electron backscatter images for specimen SnTe(x = 0)V0.04 (Fig. 6), the SEM magnification, the area of the field of view of the micrograph, along with the number, number density, and size range of the SiCNP clusters Magnification Area Number Areal Average Range Area (µm2) of SiC number diameter (nm) fraction spots density (nm) SiC (#/µm2) X 25000 17.0 12 0.71 160 54 – 560 0.025 X 50000 4.2 16 3.8 68 21 – 120 0.021 Figure 7.7. A backscattered electron image of a polished surface of specimen SnTe(x = 0)V0.04 indicates darker regions with a lower average atomic weight, likely composed primarily of the SiCNP. Using EDS, silicon was detected at the location of spectrum one, at approximately 13 at%, but not at spectrum two, with <1 at%, confirming that the dark regions contain a high concentration of the SiCNP. The specimen was osmium coated prior to imaging to reduce localized charging. 200 7.3.2 Elastic modulus as a function of temperature, matrix composition and SiCNP volume fraction For the six SnTe1+X -ySiCNP (0 < y < 4 vol %) composite specimens included in this study, the Young’s modulus (Figure 7.8) and shear modulus (not shown) decreased approximately linearly with increasing temperature over the entire temperature range from room temperature to 663 K. Equations (4a) and (4b) were fitted to the high temperature Young’s modulus, E, and shear modulus, G, data, respectively, for temperature T E = ERT (1 - bTE(T-TRT) ) (4a) G = GRT (1 - bTG(T-TRT) ) (4b) where ERT and GRT are the RT intercepts, bTE and bTG are measures of the changes in E and G, respectively, with temperature, T, and RT TRT (295 K). The values of the leastsquares parameters ERT, GRT, bTE and bTG are listed in Table 7.3 for each SnTe1+X -SiCNP specimen included in this study. The near- superposition of the heating and cooling curves for the elastic moduli (Figure 7.8) indicates the absence of either (1) significant microcracking or (2) bloating. Microcracking, if present, can reduce the elastic modulus of a material [Fan 2012; Case 1993; Chotard 2008]. Moreover, for materials that are microcracked at room temperature, heating to a homologous temperature of about 0.6 can induce microcrack healing, which in turn leads to a hysteresis in the modulus versus temperature curve [Chotard 2008; Case 1981a]. Bloating, which refers to the generation of porosity during heating, is often associated with solid phase decomposition reactions within a specimen [Ni 2011]. 201 SnTe(x=0)V0 Heating SnTe(x=0)V0 Cooling (a) 60 55 50 300 400 500 600 700 Young's Modulus (GPa) Young's Modulus (GPa) 65 65 60 55 50 300 SnTe(x=0)V0.02 Heating SnTe(x=0)V0.02 Cooling (c) 60 55 50 300 400 500 600 700 65 55 50 400 500 600 700 Young's Modulus (GPa) Young's Modulus (GPa) 60 300 600 700 60 55 50 300 400 500 600 700 Temperature (K) SnTe(x=0)V0.04 Heating SnTe(x=0)V0.04 Cooling (e) (c) 500 SnTe(x=0)V0.03 Heating SnTe(x=0)V0.03 Cooling (d) (c) Temperature (K) 65 400 Temperature (K) Young's Modulus (GPa) Young's Modulus (GPa) Temperature (K) 65 SnTe(x=0)V0.01 Heating SnTe(x=0)V0.01 Cooling (b) (f) 60 SnTe(x=0.016)V0 Heating SnTe(x=0.016)V0 Cooling 55 50 45 300 400 500 600 700 Temperature (K) Temperature (K) Figure 7.8. The Young’s modulus versus temperature for the six SnTe–SiCNP specimens included in this study. For each specimen, there is no observable hysteresis between the heating and cooling curves indicating the lack of significant microcracking or bloating over the test temperature range (room temperature to 663 K). The error bars are smaller than the plotting symbols for each modulus value in (a–f). 202 Table 7.3. For the six SnTe-SiCNP specimens included in this study, results of a least squares fit to equation (4a) for the Young’s modulus, E, versus temperature, T, data and to equation (4b) for the shear modulus, G, versus T, data. The coefficients of determination for the fit of the E versus T and the G versus T data are given by and , respectively. Specimen SnTe(x=0)V0 Vol. fraction SiC 0.00 SnTe(x=0)V0.01 0.01 SnTe(x=0)V0.02 0.02 SnTe(x=0)V0.03 0.03 SnTe(x=0)V0.04 0.04 Young's Modulus (GPa) SnTe(x=0.016)V0 0.00 ERT bTE (K-1)×10-4 R2 62.87 ± 0.13 60.84 ± 0.10 63.40 ± 0.12 64.48 ± 0.09 61.04 ± 0.07 55.31 ± 0.15 5.83 ± 0.09 5.33 ± 0.07 5.82 ± 0.08 5.38 ± 0.06 5.44 ± 0.05 5.07 ± 0.12 0.993 25.27 ± 0.05 0.995 24.54 ± 0.05 0.994 25.33 ± 0.05 0.996 25.42 ± 0.05 0.998 24.41 ± 0.03 0.995 21.70 ± 0.07 GRT bTG (K-1) R2 6.14 ± 0.09 5.62 ± 0.09 6.20 ± 0.08 5.72 ± 0.08 5.37 ± 0.05 5.29 ± 0.14 0.994 0.992 0.995 0.994 0.998 0.980 ERT bEL RT TKINK bEU Temperature (K) Figure 7.9. Schematic representation of the bilinear Young’s modulus, E, versus temperature, T, behavior showing TKINK and the associated slope change in the E versus T curve. 203 (a) 1.0 SnTe(x=0)V0 Heating SnTe(x=0)V0 Cooling 0.5 Residual (GPa) Residual (GPa) 1.0 0.0 -0.5 -1.0 (b) 0.5 0.0 -0.5 -1.0 300 400 500 600 700 300 Temperature (K) (c) 1.0 SnTe(x=0)V0.02 Heating SnTe(x=0)V0.02 Cooling 0.5 0.0 -0.5 -1.0 (d) 600 700 SnTe(x=0)V0.03 Heating SnTe(x=0)V0.03 Cooling 0.5 0.0 -0.5 400 500 600 700 300 Temperature (K) (e) 400 500 600 700 Temperature (K) 1.0 SnTe(x=0)V0.04 Heating SnTe(x=0)V0.04 Cooling 0.5 Residual (GPa) Residual (GPa) 500 -1.0 300 1.0 400 Temperature (K) Residual (GPa) Residual (GPa) 1.0 SnTe(x=0)V0.01 Heating SnTe(x=0)V0.01 Cooling 0.0 -0.5 -1.0 (f) SnTe(x=0.016)V0 Heating SnTe(x=0.016)V0 Cooling 0.5 0.0 -0.5 -1.0 300 400 500 600 700 300 400 500 600 700 Temperature (K) Temperature (K) Figure 7.10. The residuals (equation 5) representing the difference between the experimental Young’s modulus data and the modulus values predicted from the leastsquares fit of the E versus T data to equation (4). The dashed lines at 0.05 GPa are a guide, representing roughly ± 0.01 ERT, where ERT is defined as the room temperature intercept of the modulus–porosity relationship given by equation (4a). A change in the E versus T slope occurs at approximately 543 K (TKINK). 204 7.3.2.1 Bilinear model of elastic modulus versus temperature Although in this study the decrease in Young’s modulus for SnTe1+X is approximately linear with increasing T, we will consider (1) a bilinear model of the elastic modulus versus temperature behavior and (2) composition-dependent differences and subsequent elastic modulus differences between ingots SnTe(x=0) and SnTe(x=0.016) in this study. A schematic of the bilinear model for the E versus T behavior is shown in Figure 7.9. The motivation for the bilinear model comes from examining the residuals obtained from the least-squares fit of the linear model (equation (4a)) to the SnTe1+X E versus T data. For a given Young’s modulus, temperature data pair (Ej, Tj) the jth residual, residj is defined as residj = (Ej - Epred) (5) where Epred = the E value predicted by the least-squares fit to equation (4a) for T = Tj. Thus, for a given data set including a total of N data pairs (Ej, Tj), the least-squares procedure yields N residuals, which measure the differences between the data and the curve obtained by the least-squares fit. Using the Epred values obtained from the linear model, the residuals for each of the six specimens (Figures 7.10a – 7.10f) show maxima in the vicinity of 543 K. In contrast to the linear model, for the bilinear model for the E versus T behavior, equation 4a is applied separately for the two contiguous temperature segments: (i) T < TKINK and (ii) T > TKINK (Tables 7.4 and 7.5, Figure 7.9). At the temperature T = TKINK , the E versus T slope change is characterized by Δb, such that 205 Table 7.4. For two separate temperature ranges, RT – 543 K, and 543 K – 663 K, the E versus T was fit to equation (4a). Specimen Temperature range (K) SnTe(x=0)V0 RT to 543 543 to 663 SnTe(x=0)V0.01 RT to 543 543 to 663 SnTe(x=0)V0.02 RT to 543 543 to 663 SnTe(x=0)V0.03 RT to 543 543 to 663 SnTe(x=0)V0.04 RT to 543 543 to 663 SnTe(x=0.016)V0 RT to 543 543 to 663 Fitted parameters ERT bTE (×10-4 K-1) 62.6 ± 0.1 5.4 ± 0.1 66.0 ± 0.3 7.1 ± 0.2 60.7 ± 0.1 5.1 ± 0.1 61.9 ± 0.7 5.8 ± 0.3 63.2 ± 0.1 5.5 ± 0.1 65.5 ± 0.6 6.7 ± 0.2 64.3 ± 0.1 5.1 ± 0.1 66.0 ± 0.3 6.1 ± 0.1 61.0 ± 0.1 5.3 ± 0.1 62.0 ± 0.1 5.9 ± 0.1 55.0 ± 0.1 4.5 ± 0.1 56.6 ± 1.2 5.8 ± 0.6 R2 0.999 0.995 0.993 0.972 0.993 0.987 0.995 0.996 0.996 0.991 0.991 0.907 Table 7.5. Comparison of Δb, equation (6), for this study and LAST [Ren 2010]. Specimen SnTe(x=0)V0 Material Type SnTe SnTe – 1 vol% SnTe(x=0)V0.01 SiCNP SnTe – 2 vol% SnTe(x=0)V0.02 SiCNP SnTe – 3 vol% SnTe(x=0)V0.03 SiCNP SnTe – 4 vol% SnTe(x=0)V0.04 SiCNP SnTe(x=0.016)V0 SnTe1.016 155A (5K/min) Ag0.86Pb19Sb1.0Te20 155B (5K/min) Ag0.86Pb19Sb1.0Te20 155B (2K/min) Ag0.86Pb19Sb1.0Te20 TKINK 543 K 543 K Δb -0.31 -0.14 Reference This study This study 543 K -0.22 This study 543 K -0.20 This study 543 K -0.11 This study 543 K 523 K 523 K 523 K -0.27 +0.47 +0.51 +0.47 This study [Ren 2010] [Ren 2010] [Ren 2010] 206 b  bEL  bEU bEL (6) where bEL is the slope of the lower temperature region, T < TKINK and bEU is the slope of the region upper temperature segment, T > TKINK (Table 7.5 and Figure 7.9). If, as suggested above, we let TKINK = 543 K and apply equation 4a in a bilinear (piecewise) manner from TRT to TKINK and from TKINK to TMAX then the residuals for the bilinear model are reduced in magnitude (Figure 7.11) compared to the linear model and the maximum near T = 543 K largely disappears (Figure 7.10). This result indicates that the bilinear model is useful and that 543 K is a reasonable estimate for TKINK. Based on a work by Hoang et al.[Hoang 2005], Ren et al. [Ren 2010] derived a model for the similar bilinear E versus T behavior observed for a lead-antimony-silvertellurium thermoelectric material (LAST, Ag0.86 Pb19Sb1.0Te20), where a change in the slope of E, versus was attributed to an order-disorder transition. The change in slope, Δb with increasing temperature (equation (6)) was positive in the Ren et al. study [Ren 2010] while in this study Δb is negative (Table 7.5). However, the values of TKINK are similar for the two studies, namely 523 K for LAST [Ren 2010] and 543 K for this study. In terms of the possible mechanism for the bilinear behavior of SnTe in this study, it is important to note that the LAST materials studied by Ren et al. [Ren 2010] are essentially silver and tin-doped PbTe and that PbTe is quite similar to SnTe in many respects. For example, both SnTe and PbTe are IV-VI tellurides that crystallize in the rocksalt structure [Lawson 1951; Noda 1987; Beattie 1969]. Other physical properties for SnTe and PbTe that also are similar include the lattice constants, a0 (SnTe a0 = 0.6304 nm, PbTe a0 = 0.6454 nm) [Noda 1987; Mariano 1967; Seddon 1976], hardness, H (SnTe 207 1.0 SnTe(x=0)V0 Heating SnTe(x=0)V0 Cooling (a) 0.5 Residual (GPa) Residual (GPa) 1.0 0.0 -0.5 -1.0 0.5 0.0 -0.5 -1.0 300 400 500 600 700 300 Temperature (K) 1.0 SnTe(x=0)V0.02 Heating SnTe(x=0)V0.02 Cooling (c) 0.5 0.0 -0.5 -1.0 600 700 SnTe(x=0)V0.03 Heating SnTe(x=0)V0.03 Cooling (d) 0.5 0.0 -0.5 400 500 600 700 300 Temperature (K) 1.0 SnTe(x=0)V0.04 Heating SnTe(x=0)V0.04 Cooling (e) 400 500 600 700 Temperature (K) 0.5 Residual (GPa) Residual (GPa) 500 -1.0 300 1.0 400 Temperature (K) Residual (GPa) Residual (GPa) 1.0 SnTe(x=0)V0.01 Heating SnTe(x=0)V0.01 Cooling (b) 0.0 -0.5 -1.0 SnTe(x=0.016)V0 Heating SnTe(x=0.016)V0 Cooling (f) 0.5 0.0 -0.5 -1.0 300 400 500 600 700 300 400 500 600 700 Temperature (K) Temperature (K) Figure 7.11. Unlike Figure 7.10, the residuals represent the difference between the experimental data and the least-squares fit performed to equation (4) in a piecewise manner, that is least-squares fits were performed separately for the two temperature intervals (i) TRT to TKINK and (ii) TKINK to the maximum test temperature. As in Figure 7.10, the dashed lines at 0.05 GPa represent roughly ± 0.01 ERT, where ERT is defined as the room temperature intercept of the modulus-porosity relationship given by equation (4a). In (a) – (f), the vertical line at 543 K represents TKINK, the temperature at which the change in the E versus T slope occurs. 208 H = 0.78 GPa, PbTe H = 0.76 GPa) [Cui 2003], thermal expansion coefficient, α (PbTe α = 20.4 × 10-6 K-1, SnTe α = 21.3 × 10-6 K-1) [Houston 1968; Belson 1970] and melting points, TMP (PbTe TMP = 1197 K, SnTe TMP = 1079 K) [Liu 2009]. In addition, the elastic moduli of cubic materials such as SnTe and PbTe can be compared using the mean of the Hashin-Strikman (H-S) bounds. (The H-S bounds, which are computed from the single crystal elastic constants, can be used to estimate the elastic moduli of dense polycrystalline materials without crystallographic texture). For SnTe, the mean of the H–S bounds for the Young’s modulus, E, and shear modulus, G, is 55.6 and 21.7 GPa, respectively [Beattie 1969; Simmons 1971]. For PbTe, the mean of the H–S bounds Young’s moduli and shear moduli is 58.1 and 23.0 GPa, respectively [Houston 1968]. Thus, both the E and G values of SnTe and PbTe are similar. Thus, given both the similarity between SnTe and PbTe (and LAST) and the tendency of SnTe to crystallize with an off-stoichiometric composition (as will be discussed in the following section), the order-disorder mechanism proposed by Ren et al. [Ren 2010] may also apply to SnTe. However, it must be noted that in LAST, the Ag precipitates as nanoparticles of Ag2Te [Pei 2011; Cook 2009], but the order–disorder mechanism related to the modulus–temperature kink in LAST [Ren 2010] may be more closely associated with the matrix rather than the Ag2Te nanoparticles. SnTe does not have a similar nanostructure. Thus, further research is needed to explore the kink behavior in the E versus T behavior of SnTe. 209 7.3.2.2 Effects of chemical composition and porosity on elastic moduli of the SnTe1+X matrix In addition to the E versus T slope discussed in the previous section, we shall also consider the nature of changes in elastic modulus both as (i) a function of the chemical composition and (ii) as a function of residual porosity in the specimens. We shall first consider the possible effects on the elastic modulus due to compositional differences. In this study, the composition of matrix may be written as SnTe1+x, where for ingot SnTe(x=0), x = 0 and for ingot SnTe(x=0.016), x = 0.016. For specimens without SiCNP addition, the room temperature Young’s modulus of specimen SnTe(x=0)V0 from ingot SnTe(x=0) is 62.9 GPa (Table 7.2 and Figure 7.8) and the Young’s modulus of specimen SnTe(x=0.016)V0 from ingot SnTe(x=0.016) is 55.3 GPa. These differences in the elastic moduli between ingots SnTe(x=0) and SnTe(x=0.016) may be related to differences in the chemical composition between the two ingots since significant changes in Young’s modulus and bulk modulus have been observed to accompany shifts in the chemical composition of solid solution systems [Ren 2007]. As an example, for 14 solid solution systems from the literature, Ren et al. [Ren 2007] characterized composition-induced changes in Young’s modulus, E, and bulk modulus, K, in terms of Emax/Emin and Kmax/Kmin, where the subscripts “max” and “min” refer to the maximum value in E or K, respectively, over the entire compositional range of the solid solution. As reviewed by Ren et al., Emax/Emin and Kmax/Kmin values as large as 3 have been observed in the literature [Ren 2007]. To further examine the numerical values of composition-related changes in mechanical properties, Table 7.6 compares the observed changes in E with composition 210 Table 7.6. Fractional changes in Young’s modulus, E, shear modulus, G, and hardness H as a function of changes in composition from n1 to n2 [this study, [Kawaharada 2004; Ravinder 2001; Schenk 1998]. Material SnTe1+x NiZrSn1-xSbx NiZrSn1-xSbx Mn1xCdxFe2O4 Mn1xCdxFe2O4 Zn1-xCdxTe Property Property at composition x = n1 E x=0 62.9 GPa E x = 0.01 111 GPa E x = 0.01 111 GPa E x=0 352 GPa G x=0 140 GPa H x=0 700 MPa Property at composition x = n2 x = 0.016 55.3 GPa x = 0.02 89.2 GPa x = 0.05 71.8 GPa x = 0.2 256 GPa x = 0.2 108 GPa x = 0.04 930 MPa 211 Fractional References change in property -0.12 This study -0.20 -0.27 [Kawaharada 2004] [Kawaharada 2004] [Ravinder 2001] -0.23 [Ravinder 2001] +0.33 [Schenk 1998] -0.35 for SnTe1+X in this study, with additional literature data on compositionally-related changes in Young’s modulus, E, shear modulus, G, and hardness H for several tellurides, antimonides and oxides [Ravinder 2001; Kosuga 2005; Kawaharada 2004; Schenk 1998]. In Table 7.6, the compositional change in SnTe1+X comes about from the creation of Sn vacancies in the lattice [Rogacheva 1991] (Section 7.3.1) and each of the other examples in Table 7.6 is substitutional systems. From Table 7.6 and the Ren et al. study [Ren 2010], it appears to be feasible that the 10% difference in the elastic moduli observed between ingots SnTe(x=0) and SnTe(x=0.016) (Table 7.3) in this study might be attributed to compositional differences. The lattice parameter measurements for ingots SnTe(x=0) and SnTe(x=0.016) (Section 7.3.1) also are consistent with compositional differences between ingots SnTe(x=0) and SnTe(x=0.016). In addition to composition differences between the ingots SnTe(x=0) and SnTe(x=0.016), differences in the volume fraction porosity, P, among the specimens can also affect the elastic moduli of the specimens. In general, frequently used relationship between the Young’s modulus, E, shear modulus, G, and P are given by [Rice 1998] E = E0 exp(-bPEP) (7a) G = G0 exp(-bPGP) (7b) where E0 and G0 are the Young’s modulus or shear modulus at zero volume fraction porosity, P. The mean of the H-S bounds (section 7.3.2.1) can be used to estimate E0 and G0. The parameters bPE and bPG are material dependent constants [Rice 1998] that characterize the rate of decrease in the elastic moduli with increasing P. The values of bPE and bPG for SnTe1+X are not available in the literature and in fact modulus versus 212 porosity data is very limited in the TEs literature. The only two studies in the literature for which the porosity dependence of E and G for TE materials are Ni et al. [Ni 2009] study of LAST and Schmidt et al. [Schmidt 2013] study of YbAl3. For LAST, bPE = 3.5 ± 0.2 and bPG = 3.5 ± 0.2, and for YbAl3, bPE = 2.34 ± 0.06 and bPG = 2.38 ± 0.06. In order to compare the single crystal of E0 and G0 from the literature for SnTe [Beattie 1969; Simmons 1971] to the E and G of the polycrystalline specimens in this study, we must account for the effect of porosity in the polycrystalline specimens (Equations (7a), (7b)), since the single crystal data represent materials for which P = 0. Beattie’s single crystal moduli [Beattie 1969] were corrected for the P for the specimens without SiCNP additions (SnTe(x = 0)V0 and SnTe(x = 0.016)V0) using ( ) ( ) (7c) ( ) ( ) (7d) where ECORR and GCORR are the porosity-corrected values of E0 and G0. Also, for and we use the range of values obtained from the LAST and YbAl3 studies [Ni 2009; Schmidt 2013]. The experimentally measured RT values of Eexp and Gexp are compared to the corrected single crystal values, ECORR and GCORR, in Table 7.7. Although the measured Eexp and Gexp are consistently higher than these calculated values from Beattie’s single crystal data [Beattie 1969; Simmons 1971], the Eexp and Gexp values for the tellurium-rich specimen, SnTe(x = 0.016)V0, are in significantly better agreement (3–8 % higher) with the porosity-corrected Beattie data than the Eexp and Gexp values for the stoichiometric specimen, SnTe(x = 0)V0 (which are 21–26 % higher). A key question is “Were Beattie’s [Beattie 1969] single crystal SnTe specimens also Te rich?” According to Baughman [Baughman 1969], Beattie’s SnTe single crystal 213 Table 7.7. In order to compare the room temperature experimental values of Young’s modulus, Eexp, and shear modulus, Gexp, for this study’s specimens without SiCNP additions to the single crystal values of E0 and G0 values from SnTe from the literature [Beattie 1969; Simmons 1971], equations (7c) and (7d) were used to calculate porosity corrected values for the single crystal data using the modulus-porosity slope data for two thermoelectric materials, namely YbAl3 (bPE = 2.34, bPG = 2.38) and LAST (bPE = 3.5, bPG = 3.5) [Schmidt 2013; Ni 2009]. This study’s experimental Eexp and Gexp values for the tellurium rich specimen, SnTe(x=0.016)V0, are in the best agreement with the range of porosity corrected values ECORR and GCORR. Specimen SnTe (x=0)V0 SnTe (x=0.016)V0 P Eexp ECORR (GPa) (bPE=2.34) (GPa) 62.87 0.028 ± 0.13 52.03 55.31 0.022 ± 0.15 52.76 ECORR (bPE=3.5) (GPa) 50.36 51.43 214 Gexp ECORR(GPa) (bPE=2.38) (GPa) 25.27 ± 0.05 20.68 21.70 ± 0.07 20.97 ECORR(bPE=3.5) (GPa) 20.04 20.46 specimens were likely Te-rich. Beattie’s specimens were grown by the Czochralski technique, “from a stoichiometric mixture of the components” [Baughman 1969], however, the resulting SnTe crystals had “an indication that the original crucible charge was not exactly stoichiometric …” giving a tellurium-rich phase, as supported by Baughman’s observation [Baughman 1969] that “The termination of the tin telluride crystal … is a typical example of this occurrence” [Baughman 1969]. Thus, the SnTe composition is likely off stoichiometry for the available single crystal elasticity literature [Beattie 1969], so that in turn the polycrystalline aggregate values [Simmons 1971] calculated from the Beattie single crystal data in turn also represents the elastic moduli for non-stoichiometric SnTe. Unfortunately, the precise chemical composition of Beattie’s SnTe specimens is not given by either Beattie [Beattie 1969] or Baughman [Baughman 1969]. Thus this study is the first study in the literature to show the apparent sensitivity of the elastic moduli of SnTe1+X to small changes in x, although in the literature [Ren 2007] the elastic moduli of other compounds have a documented sensitivity to small changes in composition. 7.3.2.3 Comparison of the Young’s modulus versus elasticity behavior of polycrystalline SnTe1+X to the literature The values of ERT and bTE found in this study for SnTe1+X are comparable to those measured for PbTe and LAST [Ren 2010; Houston 1968], which are also chalcogenides and have the same crystal structure as SnTe (Table 7.8). If we compare the ERT and bTE values SnTe1+X -SiCNP more broadly with other thermoelectric materials, namely with data for selected skutterudites (Table 7.8) we find generally smaller values of bTE, namely 215 Table 7.8. Comparison of the Young’s modulus, E, versus temperature behavior for a variety of brittle materials, SnTe [this study], PbTe [Houston 1968; Simmons 1971; Ren 2010], and selected skutterudite TE materials [Schmidt 2012; Ravi 2008]. The parameters ERT and bRT were obtained via a least-squares fit of each of the data sets to equation (4a). The coefficient of determination, R2, was equal to or greater than 0.98 for each data set, indicating that equation (4a) describes the E versus temperature behavior well. Data Temp ERT bTE R2E Tech- References Material points Range (GPa) (× 10-4 nique (K) K-1) SnTe1+x (PC) a,b 25 b RT – 55.3 – 5.07 – 0.993 RUS This study c 663 64.5 5.83 – 0.998 PbTe (SC) a 5 100 – 58.4 5.7 0.99 IE d [Houston 303 1968; Simmons 1971]e Ag0.86Pb19 12 f RT – 61 – 64 5.9 – 0.98 RUS [Ren 2010] a c Sb1.0Te20 (PC) 523 6.6 Co0.95Pd0.05 13 295 – 140.4 1.9 0.99 RUS [Schmidt c Te0.05Sb3 573 2012] doped with 0.1 at.% Ce (PC) a Co0.95Pd0.05 13 g 295 – 128.2 – 1.6 to 0.99 RUS [Schmidt a c Te0.05Sb3 (PC) 573 138.7 1.8 2012] Ce0.9Fe3.5Co0.5 13 295 – 127.7 2.5 0.99 RUS [Schmidt c Sb12 (PC) a 573 2012] CoSb3 – doped 60 – 293 – 137.3 – 2.1 – 0.99 IE d [Ravi 2008] i a h (PC) 64 597 140.8 2.2 CeFe3RuSb12 62 – 293 – 132.1 – 2.2 – 0.99 IE d [Ravi 2008] i a j (PC) 63 597 138.1 2.3 a Single crystal denoted by SC, polycrystalline denoted by PC b Includes each of the six specimens included in this study (Table 7.1), measured with 25 data points per specimen c Resonant ultrasound spectroscopy d Impulse excitation e The Young’s modulus from Houston was obtained as the average of the Hashin and Shtrikman bounds as calculated by Simmons and Wang [Simmons 1971] f Each of 3 specimens were measured, 12 data points per specimen g Each of 2 specimens were measured, 13 data points per specimen h Each of 4 specimens were measured, 60 to 64 data points per specimen i ERT and bTE from calculation in [Schmidt 2012] j Each of 4 specimens were measured, 62 to 63 data points per specimen 216 between 1.6 and 2.3 K-1 [Schmidt 2012; Ravi 2008] (Table 7.8) although the ERT values for the skutterudites are higher than those for SnTe and PbTe. 7.3.2.4 Effect of volume fraction of SiCNP on the elastic moduli of the SnTe1+X -SiCNP composites Since the maximum surface thermal stress, max, produced by a thermal transient is function of the Young’s modulus (equation 3), in order to be able to estimate the insevice max for TE composite, one must determine how the added changes in the makeup of the composite (such as adding SiCNP) change the Young’s modulus of the material. To model the effect of the SiCNP on the Young’s modulus, EC, of the SnTe1+X SiCNP composites included in this study, where in this case Er is the elastic modulus of the SiCNP reinforcement, Em is the Young’s modulus of the matrix phase, Vm is the volume fraction of the SnTe1+X matrix and Vr is the volume fraction of the SiCNP reinforcement, we will use the following four composite models (equations 8 – 12), namely: the rule of mixtures (ROM) model EC = Vm Em + Vr Er (8) the Reuss constant strain model [Hashin 1962], V V 1  m  r EC Em Er (9) the Hashin particulate composite model [Hashin 1962; Bedolla 2012; Couturier 1997],  E V  Er Vr  1  EC  Em  m m  ErVm  Em Vr  1 (10) and the Halpin-Tsai composite model [Halpin 1992] 217  1  2( a / b )qVr   EC  Em   ErVm  Em Vr  1 (11) where EC is a function of the boundary condition parameter, q,  Er    1 Em   q  Er   a     2   Em   b  (12) and a/b is the length/thickness ratio of the reinforcing phase. Note that each of the composite models above (equations (8)–(12)) is a function of Vr, the volume fraction of the reinforcing phase. The particle size of the reinforcing phase is not explicitly included in any of the models, although it is assumed that the dimensions of the reinforcing particles are much smaller than the specimen dimensions. For the calculations of EC from equations (8) – (12), the volume fractions of SiCNP, Vr, in this study were 0.0, 0.01, 0.02, 0.03 and 0.04. Since the SiCNP are approximately equiaxed, the particle aspect ratio, a/b, in the Halpin-Tsai equation was set to unity. The value of the Young’s modulus for SnTe1+X matrix, Em, was set at 67.1 GPa, which is the RT E value measured on specimen SnTe(x=0)V0 (without SiCNP addition) and then corrected to zero porosity (Section 7.3.2.2). The Young’s modulus of SiCNP, the reinforcing phase, Er, was set to 450 GPa, which is the Young’s modulus for bulk SiC [51 52]. Figure 7.12 includes the experimental values of the Young’s modulus of the SnTe1+X - SiCNP, the experimental values corrected to zero porosity and the predictions of the four composite models (equations (8) – (12)). The porosity, P, of the specimens included in this study ranged over a relatively narrow range, from 0.022 to 0.051 (Table 7.1). However, each composite specimen had a 218 Young's modulus (GPa) 80 Rule of Mixtures Halpin-Tsai, a/b = 1 Measured Adjusted to P = 0 by bPE = 2.34 Hashin model Adjusted to P = 0 by bPE = 3.5 Reuss model 75 70 65 60 0.00 0.01 0.02 0.03 0.04 Volume fraction of SiCNP Figure 7.12. Comparison of the experimental Young’s modulus, E, results for the composites SnTe1+X –ySiC (y = 0.0 to 0.04) with the four composite models given by equations (8) – (12). Also plotted are the values of E corrected to zero porosity using equation (7a). 219 porosity that was different than the other specimens (Table 7.1), thus in order to compare the experimental data with the four elasticity models, the Young’s modulus was corrected to correspond to zero porosity. Figure 7.12 shows both (i) the experimental Young’s modulus data and (ii) the data corrected to zero P (using equation 7a). The Reuss model (equation (9)) and the Hashin model (equation (10)) correspond best with the experimental data. 7.5 Summary and conclusions The linear decrease in Young’s modulus, E, (Figure 7.8 and Table 7.3) and shear modulus, G, (Table 7.3) with increasing temperature is relatively described well by equations 4(a) and 4(b), respectively. The lack of significant hysteresis between the heating and cooling curves for E and G indicates that no significant microcracking or bloating occurred over the test temperature range (RT to 663 K). However, there appears to be a bilinear behavior for the elastic modulus versus temperature behavior (Figure 7.9), with a change in the modulus versus temperature behavior at about 543 K. This behavior might be explained by an order–disorder mechanism similar to that proposed by Ren et al. [Ren 2010]. The addition of up to 4 vol% SiCNP to the SnTe1+X matrix lead to relatively modest changes in (1) the observed modulus versus temperature slope and (2) the RT elastic moduli, ERT. Changes in ERT can be described relatively well by either the Reuss [Hashin 1962]or the Hashin [Hashin 1962] composite models (Figure 7.12). In the study, the elastic modulus changes due to (i) the possible order–disorder transition in offstoichiometric SnTe and (ii) the addition of up to 4 vol% SiCNP was small. However compositional shifts in the SnTe matrix seem to represent a much larger source of 220 perturbation on the elastic moduli of the SnTe1+X – SiCNP composite materials than either the SiCNP additions or the bilinear E versus T behavior. For the SnTe1+X matrix, in the absence of SiCNP addition, the compositional differences between SnTe(x = 0) and SnTe(x = 0.016) may be responsible for the differences in the measured elastic moduli between the two ingots. Similar significant changes in elastic moduli with composition of solid solution systems have been reported in the literature for other materials (Table 7.6), but this study is the first to report the sensitivity of the elastic moduli to compositional changes in SnTe. Future studies should explore functional relationship of the elastic moduli of SnTe1+X and the Te composition, especially since there are some questions about the stoichiometry of the SnTe elasticity data currently available in the literature. ACKNOWLEDGEMENTS The authors acknowledge the financial support of the Department of Energy, Revolutionary Materials for Solid State Energy Conversion Center, an Energy Frontiers Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under award number DE-SC0001054. The authors also acknowledge the use of the equipment high-temperature RUS measurements through the Oak Ridge National Laboratory’s High Temperature Materials Laboratory User Program, which is sponsored by the U. S. Department of Energy, Office of Energy Efficiency and Renewable Energy, Vehicle Technologies Program. 221 REFERENCES 222 REFERENCES [Baughman, Lefever 1969] Baughman, R., Lefever, R. Czochralski encapsulation growth of GeTe, SnTe and PbTe single crystals. Materials Research Bulletin 4 (1969) 721– 726. [Beattie 1969] Beattie, A.G. Temperature Dependence of the Elastic Constants of Tin Telluride. Journal of Applied Physics 40 (1969) 4818. [Bedolla, Lemus-Ruiz, Contreras 2012] Bedolla, E., Lemus-Ruiz, J., Contreras, A. Synthesis and characterization of Mg-AZ91/AlN composites. Materials and Design 38 (2012) 91–98. [Belson 1970] Belson, H.S. Thermal Expansion of Tin Telluride. Journal of Applied Physics 41 (1970) 422. [Bux, Fleurial, Kaner 2010] Bux, S.K., Fleurial, J.-P., Kaner, R.B. Nanostructured materials for thermoelectric applications. Chemical communications (Cambridge, England) 46 (2010) 8311–24. [Case, Kim 1993] Case, E.D., Kim, Y. The effect of surface-limited microcracks on the effective Young’s modulus of ceramics. Journal of Materials Science 28 (1993) 1885–1900. [Case, Smyth, Hunter 1981a] Case, E.D., Smyth, J.R., Hunter, O. Microcracking in Large-grain AI203. 51 (1981) 175–179. [Case, Smyth, Monthei 1981b] Case, E.D., Smyth, J.R., Monthei, V. Grain-Size Determinations. Journal of the American Ceramic Society 64 (1981) C–24–C–25. [Chotard, Soro, Lemercier, Huger, Gault 2008] Chotard, T., Soro, J., Lemercier, H., Huger, M., Gault, C. High temperature characterisation of cordierite–mullite refractory by ultrasonic means. Journal of the European Ceramic Society 28 (2008) 2129–2135. [Cook, Kramer, Harringa, Han, Chung, Kanatzidis 2009] Cook, B. a., Kramer, M.J., Harringa, J.L., Han, M.-K., Chung, D.-Y., Kanatzidis, M.G. Analysis of Nanostructuring in High Figure-of-Merit Ag1- x Pb m SbTe2+ m Thermoelectric Materials. Advanced Functional Materials 19 (2009) 1254–1259. [Couturier, Ducret, Merle, Disson, Joubert 1997] Couturier, R., Ducret, D., Merle, P., Disson, J.P., Joubert, P. Elaboration and characterization of a metal matrix composite: Al/AlN. Journal of the European Ceramic Society 17 (1997) 1861–1866. 223 [Cui, Qian, Zhao 2003] Cui, J., Qian, X., Zhao, X. Mechanical and transport properties of pseudo-binary alloys (PbTe) 1− x–(SnTe) x by pressureless sintering. Journal of alloys and compounds 358 (2003) 228–234. [D’Angelo, Case, Matchanov, Wu, Hogan, Barnard, Cauchy, Hendricks, Kanatzidis 2011] D’Angelo, J., Case, E.D., Matchanov, N., Wu, C.-I., Hogan, T.P., Barnard, J., Cauchy, C., Hendricks, T., Kanatzidis, M.G. Electrical, Thermal, and Mechanical Characterization of Novel Segmented-Leg Thermoelectric Modules. Journal of Electronic Materials 40 (2011) 2051–2062. [Fan, Case, Baumann 2012] Fan, X., Case, E.D., Baumann, M.J. The effect of indentation-induced microcracks on the elastic modulus of hydroxyapatite. Journal of Materials Science 47 (2012) 6333–6345. [Halpin 1992] Halpin, J.C. Primer on composite materials analysis. Technomic Publishing Company, Inc., Lancaster, Pennsylvania, 1992. [Hashin 1962] Hashin, Z. The Elastic Moduli of Heterogeneous Materials. Journal of Applied Mechanics 29 (1962) 143–150. [Hoang, Desai, Mahanti 2005] Hoang, K., Desai, K., Mahanti, S. Charge ordering and self-assembled nanostructures in a fcc Coulomb lattice gas. Physical Review B 72 (2005) 064102. [Houston, Strakna, Belson 1968] Houston, B., Strakna, R.E., Belson, H.S. Elastic Constants, Thermal Expansion, and Debye Temperature of Lead Telluride. Journal of Applied Physics 39 (1968) 3913. [Kaliakin 2002] Kaliakin, V.N. Introduction to approximate solution techniques, numerical modeling, and finite element methods. Marcel Dekker, New York, 2002. [Kawaharada, Kurosaki, Muta, Uno, Yamanaka 2004] Kawaharada, Y., Kurosaki, K., Muta, H., Uno, M., Yamanaka, S. Thermophysical properties of NiZrSn1−xSbx half-Heusler compounds. Journal of Alloys and Compounds 381 (2004) 9–11. [Kosuga, Uno, Kurosaki, Yamanaka 2005] Kosuga, A., Uno, M., Kurosaki, K., Yamanaka, S. Thermoelectric properties of Ag1−xPb18SbTe20 (x = 0, 0.1, 0.3). Journal of Alloys and Compounds 387 (2005) 52–55. [Kreith 1973] Kreith, F. Principles of heat transfer. Intext Educational Publishers, New York, 1973. [Lawson 1951] Lawson, W.D. A Method of Growing Single Crystals of Lead Telluride and Lead Selenide. Journal of Applied Physics 22 (1951) 1444. 224 [Liu, Liang, Zhang 2009] Liu, Y., Liang, D., Zhang, L. Thermodynamic Descriptions for the Sn-Te and Pb-Sn-Te Systems. Journal of Electronic Materials 39 (2009) 246– 257. [Lu, Fleck 1998] Lu, T.J., Fleck, N. a. The thermal shock resistance of solids. Acta Materialia 46 (1998) 4755–4768. [Manson 1966] Manson, S.S. Thermal stress and low cycle fatigue. McGraw-Hill, New York, 1966. [Mariano 1967] Mariano, a. N. Polymorphism in Some Iv-Vi Compounds Induced By High Pressure and Thin-Film Epitaxial Growth. Applied Physics Letters 10 (1967) 282. [Martin, Carey 1973] Martin, H.C., Carey, G.F. Introduction to finite element analysis. McGraw-Hill, New York, 1973. [Ni, Ren, Case, Timm 2009] Ni, J., Ren, F., Case, E., Timm, E. Porosity dependence of elastic moduli in LAST (lead–antimony–silver–tellurium) thermoelectric materials. Materials Chemistry and Physics(2009) . [Ni, Case, Stewart, Wu, Hogan, Kanatzidis 2011] Ni, J.E., Case, E.D., Stewart, R., Wu, C.-I., Hogan, T.P., Kanatzidis, M.G. Bloating in (Pb0.95Sn0.05Te)0.92(PbS)0.080.055%PbI2 Thermoelectric Specimens as a Result of Processing Conditions. Journal of Electronic Materials 41 (2011) 1153–1158. [Noda, Masumoto, Ohba, Saito, Toriumi, Iwata, Shibuya 1987] Noda, Y., Masumoto, K., Ohba, S., Saito, Y., Toriumi, K., Iwata, Y., Shibuya, I. Temperature Dependence of Atomic Thermal Parameters of Lead Chaleogenides, PbS, PbSe and PbTe. Acta Crystallographica C43 (1987) 1443–1445. [Pei, Lensch-falk, Toberer, Medlin, Snyder 2011] Pei, Y., Lensch-falk, J., Toberer, E.S., Medlin, D.L., Snyder, G.J. High Thermoelectric Performance in PbTe Due to Large Nanoscale Ag 2 Te Precipitates and La Doping. (2011) 241–249. [Pilchak, Ren, Case, Timm, Schock, Wu, Hogan 2007] Pilchak, A.L., Ren, F., Case, E.D., Timm, E.J., Schock, H.J., Wu, C.-I., Hogan, T.P. Characterization of dry milled powders of LAST (lead–antimony–silver–tellurium) thermoelectric material. Philosophical Magazine 87 (2007) 4567–4591. [Ravi, Firdosy, Caillat, Lerch, Calamino, Pawlik, Nathal, Sechrist, Buchhalte, Nutt 2008] Ravi, V., Firdosy, S., Caillat, T., Lerch, B., Calamino, A., Pawlik, R., Nathal, M., Sechrist, A., Buchhalte, J., Nutt, S. Mechanical Properties of Thermoelectric Skutterudites. Space Technology and Applications International Forum—STAIF 2008(2008) 656–662. 225 [Ravinder, Alivelumanga 2001] Ravinder, D., Alivelumanga, T. Room temperature elastic behaviour of cadmium substituted manganese ferrites. Materials Letters(2001) 1–6. [Ren, Case, Morrison, Tafesse, Baumann 2009a] Ren, F., Case, E.D., Morrison, A., Tafesse, M., Baumann, M.J. Resonant ultrasound spectroscopy measurement of Young’s modulus, shear modulus and Poisson's ratio as a function of porosity for alumina and hydroxyapatite. Philosophical Magazine 89 (2009) 1163–1182. [Ren, Case, Ni, Timm, Lara-Curzio, Trejo, Lin, Kanatzidis 2009b] Ren, F., Case, E.D., Ni, J.E., Timm, E.J., Lara-Curzio, E., Trejo, R.M., Lin, C.-H., Kanatzidis, M.G. Temperature-dependent elastic moduli of lead telluride-based thermoelectric materials. Philosophical Magazine 89 (2009) 143–167. [Ren, Case, Sootsman, Kanatzidis, Kong, Uher, Lara-Curzio, Trejo 2008] Ren, F., Case, E.D., Sootsman, J.R., Kanatzidis, M.G., Kong, H., Uher, C., Lara-Curzio, E., Trejo, R.M. The high-temperature elastic moduli of polycrystalline PbTe measured by resonant ultrasound spectroscopy. Acta Materialia 56 (2008) 5954–5963. [Ren, Case, Timm, Lara-Curzio, Trejo 2010] Ren, F., Case, E.D., Timm, E.J., LaraCurzio, E., Trejo, R.M. Anomalous temperature-dependent Young’s modulus of a cast LAST (Pb–Sb–Ag–Te) thermoelectric material. Acta Materialia 58 (2010) 31– 38. [Ren, Case, Timm, Schock 2007] Ren, F., Case, E.D., Timm, E.J., Schock, H.J. Young’s modulus as a function of composition for an n-type lead–antimony–silver–telluride (LAST) thermoelectric material. Philosophical Magazine 87 (2007) 4907–4934. [Rice 1998] Rice, R.W. Porosity of Ceramics. Marcel Dekker, New York, 1998. [Rogacheva, Gorne, Zhigareva, Ivanova 1991] Rogacheva, E.I., Gorne, G. V, Zhigareva, N.K., Ivanova, A.B. HOMOGENEITY REGION OF TIN MONOTELLURIDE. INORGANIC MATERIALS 27 (1991) 194–197. [Schenk, Dunog 1998] Schenk, M., Dunog, L.T.H. Solid-solution hardening of Cd 1 − x Zn x Te bulk crystals. Semiconductor Science and Technology 16 (1998) 335–339. [Schmidt, Case, Lehr, Morelli 2013] Schmidt, R.D., Case, E.D., Lehr, G.J., Morelli, D.T. Room temperature mechanical properties of polycrystalline YbAl3, a promising low temperature thermoelectric material. Intermetallics 35 (2013) 15–24. [Schmidt, Case, Ni, Sakamoto, Trejo, Lara-Curzio, Payzant, Kirkham, Peascoe-Meisner 2012] Schmidt, R.D., Case, E.D., Ni, J.E., Sakamoto, J.S., Trejo, R.M., Lara-Curzio, E., Payzant, E.A., Kirkham, M.J., Peascoe-Meisner, R.A. The temperature dependence of thermal expansion for p-type Ce 0.9 Fe 3.5 Co 0.5 Sb 12 and n-type 226 Co 0.95 Pd 0.05 Te 0.05 Sb 3 skutterudite thermoelectric materials. Philosophical Magazine 92 (2012) 1261–1286. [Seddon, Gupta, Isci, Saunders 1976] Seddon, T., Gupta, S.C., Isci, C., Saunders, G.A. The elastic constants of IV?VI compound alloys. Journal of Materials Science 11 (1976) 1756–1759. [Simmons, Wang 1971] Simmons, G., Wang, H. Single Crystal Elastic Constants and Calculated Aggregate Properties: A Handbook. The M.I.T. Press, Cambridge, Massachusetts, 1971. [Srinivas Rao, Ravinder 2003] Srinivas Rao, S., Ravinder, D. Composition dependence of elastic moduli of gadolinium-substituted nickel–zinc ferrites. Materials Letters 57 (2003) 3802–3804. [Wyckoff 1963] Wyckoff, R. Rocksalt crystal structures. Interscience, New York, 1963. [Zhao, Lu, Fleck 2000] Zhao, L.G., Lu, T.J., Fleck, N. a. Crack channelling and spalling in a plate due to thermal shock loading. Journal of the Mechanics and Physics of Solids 48 (2000) 867–897. 227 8 Mechanical properties of lower-cost, earth-abundant chalcogenide thermoelectric materials, PbSe and PbS, with additions of 0 to 4% CdS or ZnS Robert D. Schmidt 1, Eldon D. Case 1, Li-Dong Zhao 2, Mercouri G. Kanatzidis 2 1 Chemical Engineering and Materials Science Department, Michigan State University, East Lansing, MI, 48824 2 Department of Chemistry, Northwestern University, Evanston, IL, 60208 To be submitted to: Journal of Materials Science Abstract The thermoelectric (TE) properties of PbTe-based materials have been optimized to the extent that the maximum figure of merit, ZT, achieved is greater than 2. However, the tellurium content limits the application potential due to both availability and cost. Replacing the tellurium with selenium or sulfur produces an analogous thermoelectric material using strategies learned from PbTe-based materials to achieve ZT values of 1.3 to 1.6. In order to effectively incorporate these new materials into TE devices, it is important to understand materials’ response to thermally and mechanically imposed loads, which in turn requires knowledge of the mechanical properties. In this study, the hardness was determined by Vickers indentation and elastic modulus and Poisson’s ratio were measured using resonant ultrasound spectroscopy (RUS) on PbSe- and PbS-based thermoelectric specimens as a function the addition of 0-4 at% of CdS or ZnS. The hardness and moduli of PbSe-based or PbS-based TE materials are not strong functions of the addition of CdS or ZnS. With 2.0 at% or 2.5 at% Na doping, the hardness of PbSebased or PbS-based TE materials increased by about 30% and the elastic moduli decreased by 5-10%. In addition, PbS may be effectively sintered at 723 K when doped with 2.5 at% Na, but requires a higher sintering temperature when undoped. 228 Keywords: thermoelectric; hardness; elastic modulus 8.1 Introduction The figure of merit, ZT, of bulk thermoelectric (TE) materials is defined as [Tritt 2011] (1) where S is Seebeck coefficient, σ is the electrical conductivity, T is temperature, and κ is thermal conductivity. The ZT has been increased using a variety of techniques, including atomic scale doping or substitution [Gelbstein 2008; Zhu 2009; Zhao 2012b], creation of nanostructured bulk materials [Toprak 2004; Mi 2008; Hsu 2004; Zhu 2009], incorporation of nanoparticles [Androulakis 2007; Zhao 2012a; Zhou 2008; Alleno 2009; Ji 2007], or a combination of atomic, nanostructural, and microstructural length scale modifications [Biswas 2012]. The application of these techniques has been particularly successful with lead telluride-based thermoelectrics, with ZT values near or greater than 2 reported [Zhao 2013; Biswas 2012; Heremans 2008]. A limitation on the application of these materials is the use of tellurium, due to the relative scarcity of the element [Haxel 2002], but sulfur or selenium may be substituted for tellurium to produce analogous compounds [Zhao 2011; Zhao 2012a; Zhao 2012b; Zhao 2013]. The thermoelectric properties of lead sulfide and lead selenide have been optimized with the same techniques as the lead telluride-based compounds, such that a ZT value of 1.6 has been achieved for lead selenide [Zhao 2013]and 1.3 for lead sulfide [Zhao 2012a; Zhao 2012b]. Changing the chemistry of the thermoelectric however, can dramatically alter the mechanical properties 229 of the material, as has been observed in lead telluride-based materials [Ren 2007; Ren 2008]. The elastic properties of the TE materials are required as well to determine the material response to stresses, just as the transport properties are required to determine the ZT of the material. In waste heat harvesting applications, the stresses that the TE material must survive are not trivial, as TE materials are intended for an environment with thermal gradients, thermal shocks, attached to materials with thermal expansion mismatches, and subjected to external loads. In addition, a TE module is typically composed of tens or hundreds of legs of TE material electrically connected in series and thermally in parallel. This arrangement subjects each leg individually to the full thermal gradient and associated stresses, but the failure of just one leg results in the electrical failure of an entire TE module. The mechanical properties of many undoped TE materials have been published, including PbTe [Houston 1968], Mg2Si [Whitten 1965; Milekhine 2002], SnTe [Beattie 1969] and CdSe [Cline 1967], but the mechanical properties may be significantly changed when the TE material is optimized for improved ZT. The hardness and/or elastic moduli of a material have been shown to change by the addition of nanoparticles [Zhao 2008; Kvetková 2013], creation of nanoprecipitates in the bulk material [Ni 2010], doping [Gelbstein 2008], or alloying [Gelbstein 2008; Darrow 1969].Therefore, to understand the mechanical response of a TE material that may be incorporated into a device, the mechanical properties need to be measured on a TE material with the addition of dopants, with nanoparticle or nanoprecipitates. 230 8.2 Experimental procedure 8.2.1 Materials and Specimen Preparation Two tellurium-free thermoelectric materials with analogous crystal structure to PbTe-based TE materials were included in this study, using methods similar to those employed to optimize the PbTe-based TE materials [Zhao 2012b; Zhao 2013]. Ingots were fabricated from elemental materials (99.99+% purity) with nominal compositions of PbS or PbSe with x% CdS/ZnS, where the at fraction x = 1, 2, 3 and 4. For PbS, 2.5 at% Na dopant was added, and for PbSe, 2.0 at% Na dopant was added [Zhao 2012b; Zhao 2013]. In addition, ingots of PbS and PbSe without Na dopant, and an ingot of PbSe with 2.0 at% Na dopant were made. Carbon-coated fused silica tubes with the elemental material were evacuated to approximately 10-4 torr before being flame-sealed [Zhao 2012b; Zhao 2013]. The tubes were heated to 723 K in 12 h, then to 1423 K in 7 h, thermally soaked for 6 h at 1423 K and then water quenched to room temperature [Zhao 2012b; Zhao 2013]. The ingots produced were crushed into powders, sieved through a 53 μm sieve, then densified by Pulsed Electric Current Sintering (PECS, Thermal Technology SPS-10-4 or SPS-DR 2050) at 723 K (PbS), 873 K (undoped PbS) or 823 K (PbSe) for 10 min in a 20 mm diameter graphite die under 60 MPa axial pressure in an argon atmosphere [Zhao 2012b; Zhao 2013]. Except for the undoped PbS sintered at 723 K, this produced highly dense disk-shaped billets 20 mm in diameter and 9 mm thick, with volume fraction porosity less than 0.05. Using a low speed diamond saw, the billets were cut into bars, ~ 2 mm × 3 mm in cross section and approximately 9-18 mm long [Zhao 2012b; Zhao 2013]. 231 8.2.2 Elasticity Measurements In this study, the elastic moduli were determined by a standing acoustic wave technique, resonant ultrasound spectroscopy (RUS). The elastic moduli of each specimen are a function of the specimen geometry, dimensions, density and resonant frequencies. Each of the specimens in this study was of rectangular bar geometry. The dimensions of each specimen were measured by micrometers (293-832, Mitutoyo, Japan), and the mass by electronic balance (Adventurer AR2140, OHAUS, Pine Brook IL). The resonant frequencies were measured by RUS (RUSpec, Quasar International, Albuquerque, NM), in which the specimen was placed so that the corners of the specimen contacted two transducers along the specimen’s body diagonal. One transducer was swept across a range of frequencies from 10 kHz to 760 kHz in 29,999 steps and the mechanical response to the frequency was recorded in the second transducer. Each sharp peak in second transducer’s response to the driving signal (Figure 8.1) represents a mechanical resonance frequency of the specimen. The set of recorded resonant frequencies were fit to a model by commercial software (RPModel version 2.68b, Quasar International) to determine the elastic moduli. Details of the RUS procedure may be found elsewhere [Ren 2009; Ni 2010; Schmidt 2010; Migliori 1997; Schmidt 2013a]. The elastic moduli of other thermoelectric materials have been measured by RUS, including PbTe-based compounds [Ni 2013; Ni 2009a; Ren 2009; Ni 2011; Ren 2010; Morrison 2012], SnTe [Schmidt 2013b], tetrahedrite [Fan 2013], YbAl3 [Schmidt 2013a], Mg2Si [Schmidt 2012a], skutterudite [Rogl 2011; Zhang 2010; Schmidt 2012b; Schmidt 2010; Salvador 2009; Möchel 2011] and Zn4Sb3 [Bhattacharya 2006]. 232 Intensity (A.U.) 0 100 200 300 400 500 Frequency (kHz) Figure 8.1. RUS scan of PbSe specimen 233 600 8.2.3 Hardness and toughness measurements Hardness was measured by Vickers indentation at 1.98 N load on a microindenter (HMV-2000, Shimadzu, Okinawa, Japan), using the equation H  1.8544 P 2a 2 (2) where P is the indentation load and 2a is the length of the indentation diagonal [Wachtman 2009]. The Vickers hardness values from indentation of a standard block (Yamamoto Scientific Tools Lab Co. LTD, Chiba, Japan) were used to determine a calibration constant, β, of 0.95. 8.2.4 Microscopy and X-ray diffraction The PbS and PbSe specimens were examined by scanning electron microscope (SEM) on both polished and fractured specimen surfaces at 15 kV accelerating voltage and 15 mm working distance (6610LV SEM, JEOL, Tokyo, Japan). Polished surfaces were examined by secondary electron images and backscatter images to examine indentation impressions and the morphology of secondary phases with different average atomic weight. Grain size analysis was performed by the lineal intercept method on secondary electron images of fractured specimen surfaces, with a minimum of 200 intercepts per image and a stereographic projection factor of 1.5. Energy dispersive x-ray spectroscopy (Oxford EDS) was performed in the SEM at 15 kV accelerating voltage and 10 mm working distance for elemental analysis. Fracture surfaces of each specimen were examined for mode mixity, M, defined here as M = ATRANS/ATOTAL (3) 234 where ATRANS is the area of transgranular fracture, and ATOTAL is the overall area of a fractured surface. X-ray diffraction (XRD) was performed at the Michigan State University Center for Crystallographic Research with Cu Kα radiation and a step size of 0.008° 2θ (Bruker Davinci Diffractometer) from 25° to 75° 2θ on a rotary stage. Two XRD scans were performed on the sintered PbS material, (i) on the bulk densified specimen, and (ii) on powder ground by hand from the bulk specimen in a mortar and pestle. 8.3 Results and Discussion 8.3.1 Microstructural analysis Only isolated spherical porosity was observed which is consistent with dense specimens. Except for 1 at% ZnS in PbSe (with precipitates up to 4 µm), the addition of ZnS in PbS or PbSe resulted in ZnS precipitates with diameters up to 10 to 15 µm (Tables 8.1 and 8.2), observed primarily at the grain boundaries (Figures 8.2 – 8.4). The addition of 1 at% to 4 at% CdS in PbS or PbSe resulted in precipitates with diameters of up to 1 µm to 4 µm (Tables 8.1 and 8.2), observed primarily embedded in the matrix material (Figures 8.2 – 8.4). In backscatter SEM imaging, both the CdS and ZnS precipitates were uniformly distributed through the specimen matrix, with up to 15 µm ZnS precipitates observed in both the PbSe and PbS (Figure 8.4, Tables 8.1 and 8.2). No precipitates were observed in the samples of either PbS or PbSe without CdS or ZnS additions. All the precipitates observed in this study were faceted (Figures 8.4 and 8.5). In particular, the precipitate morphology was rod or plate-like for the CdS additions in both PbS and PbSe 235 Table 8.1. Composition, theoretical density, theo, measured density, meas, volume fraction porosity, P, average grain size, , typical observed inclusion size range, Incl, and mode mixity, M (equation 3) for the PbS-based specimens included in this study. Composition theoa meas P (µm) 5.9b 1.8 3.2 Incl (µm) NA NA 0.2 to 1 M PbS 7.5970 7.535 0.008 0.91 Pb0.975Na0.025S 7.5970 7.325 0.036 0.15 Pb0.975Na0.025S 7.5716 7.278 0.039 0.94 + 1 at% CdS Pb0.975Na0.025S 7.5462 7.315 0.031 5.4 0.2 to 1.5 0.98 + 2 at% CdS Pb0.975Na0.025S 7.5207 7.148 0.050 4.5 0.2 to 2c 0.92 + 3 at% CdS Pb0.975Na0.025S 7.4952 7.148 0.046 3.6 0.2 to 3c 0.93 + 4 at% CdS Pb0.975Na0.025S 7.5704 7.231 0.045 2.1 0.2 to 12 0.34 + 1 at% ZnS Pb0.975Na0.025S 7.5437 7.264 0.037 2.2 0.2 to 0.72 d + 2 at% ZnS 10 Pb0.975Na0.025S 7.5168 7.241 0.037 2.3 0.2 to 0.91 d + 3 at% ZnS 15d Pb0.975Na0.025S 7.4898 7.245 0.033 2.8 0.2 to 10 0.76 + 4 at% ZnS NA: Not applicable. No inclusions because no addition of CdS or ZnS a The theo values are calculated in Table S1 from [Zhao 2012a] b The includes a distribution of several large grains, 5 µm to 30 µm, and smaller grains, 0.2 µm to 2 µm c Inclusions were typically rod- or plate-like shape d Larger inclusions were typically stacked plate-like shapes 236 Table 8.2. Composition, theoretical density, theo, measured density, meas, volume fraction porosity, P, average grain size, , typical observed inclusion size range, Incl, and mode mixity, M (equation 3) for the PbSe-based specimens included in this study. Composition theoa meas P (µm) 9.9 4.0 3.2 Incl (µm) NA NA 0.2 to 3 PbSe 8.258 8.125 0.016 Pb0.98Na0.02Se 8.258 7.955 0.037 Pb0.98Na0.02Se 8.2243 8.112 0.014 + 1 at% CdS Pb0.98Na0.02Se 8.1906 8.048 0.017 3.6 0.2 to 2 + 2 at% CdS Pb0.98Na0.02Se 8.157 7.950 0.025 4.0 0.2 to 3b + 3 at% CdS Pb0.98Na0.02Se 8.1232 7.966 0.019 18.7 0.2 to 4b + 4 at% CdS Pb0.98Na0.02Se 8.2163 8.074 0.017 3.3 0.2 to 4 + 1 at% ZnS Pb0.98Na0.02Se 8.1746 8.005 0.021 3.2 0.2 to 10 + 2 at% ZnS Pb0.98Na0.02Se 8.1329 8.049 0.010 3.2 0.2 to 12 + 3 at% ZnS Pb0.98Na0.02Se 8.0912 7.880 0.026 7.0 0.2 to 15 + 4 at% ZnS NA: Not applicable. No inclusions because no addition of CdS or ZnS a The theo values are calculated in Table S1 from [Zhao 2013] b Inclusions were typically rod- or plate-like shape 237 M 0.07 0.28 0.17 0.92 0.81 0.98 0.27 0.73 0.83 0.51 Figure 8.2. Fracture surfaces of PbS specimens indicate primarily transgranular fracture in undoped PbS (a) changing to intergranular fracture for the Na-doped PbS specimen (b). With 1% to 4% addition of CdS, the fracture mode changes to primarily transgranular fracture (b, d), and with 1% to 4% addition of ZnS, the fracture mode changes to mixed, with majority transgranular and minority intergranular (c, e). Note the bright areas in images (c) and (e) are the ZnS precipitates. 238 Figure 8.3. Fracture surfaces of PbSe specimens indicate primarily intergranular fracture for the pure PbSe specimen, the Na-doped PbSe specimen, and the specimens with 1% addition of CdS or ZnS (a-c). With 2% to 4% addition of CdS or ZnS, the fracture mode changes to primarily transgranular fracture (d-f), although significant intergranular fracture was also observed in the specimen with 4% ZnS (f). Note the bright areas in images (d) and (f) are the ZnS precipitates. 239 Figure 8.4. Secondary images of the polished specimens show isolated, micron scale spherical porosity and minor scratches from polishing. In the specimens with 4% CdS or ZnS, precipitates of sub-micron up to approximately 15 μm were observed, particularly in backscatter electron (BSE) mode. The images of polished PbS in both secondary electron (SE) and BSE mode (a and b) only exhibit spherical pores ~1 µm, with no evidence of precipitates, consistent with the material having no additions of CdS or ZnS. 240 Figure 8.4 (cont’d) 241 Figure 8.5. Observed in backscatter, precipitates of CdS were up to ~4 μm, with some micron scale precipitates with geometry of rods or plates and sharp facets consistent with crystallographic alignment. Figure 8.6. Some ZnS inclusions in PbS resembled stacked plates. 242 (Figure 8.5). In addition, the ZnS precipitate morphology included stacked plates, particularly in the PbS matrix (Figures 8.4 and 8.6). The fracture mode changed from primarily intergranular fracture in Na-doped PbS and PbSe to primarily transgranular fracture when additions of up to 4 at% CdS or ZnS were added to either Na-doped PbS (Table 8.1, Figure 8.2) or PbSe (Table 8.2, Figure 8.3). Mode mixity, M, was examined to determine the fraction of transgranular fracture, with M = 0 indicating pure intergranular fracture and M = 1 indicating pure transgranular fracture. Undoped PbS exhibited primarily transgranular fracture, M = 0.91, while the Na-doped PbS changed to primarily intergranular, M = 0.14. The PbSe specimens without CdS or ZnS exhibited primarily intergranular fracture, with both Na doped and pure PbSe specimen M values ranging from 0.07 to 0.28. M increased to 0.51 to 0.98 for PbS and PbSe specimens with additions of 2 at% to 4 at% CdS or ZnS, as compared to the doped PbS and PbSe specimens. Generally, M was higher for specimens with additions of CdS than with additions of ZnS (Tables 8.1 and 8.2). In both the PbS with ZnS (Table 8.1) and the PbSe with ZnS (Table 8.2), M reaches a maximum at 3 at% ZnS. In contrast, there was no observed maximum in M for CdS additions to PbS and PbSe. EDS analysis of the polished surface of PbS indicated a deficiency of S (Table 8.3). The lattice parameter from XRD of the PbS specimen was 5.941 Å for the bulk, as-sintered specimen (Figure 8.7). The measured lattice parameters for sintered PbS are larger than the lattice parameter of 5.932 Å for the same PbS material prior to sintering [Zhao 2012a], and consistent with a change in the concentration of S, as indicated by EDS analysis. 243 Table 8.3. EDS results from four area scans of a polished PbS specimen, indicating a higher concentration of Pb than S. Location ID Pb 50.4 50.4 50.9 50.5 50.8 Intensity (AU) 1 2 3 4 5 Atomic % S 49.6 49.6 49.1 49.5 49.2 25 50 75 Position (2) Figure 8.7. XRD pattern of the sintered PbS specimen. 244 The undoped PbS specimen required a higher sintering temperature (873 K) than the Na-doped PbS specimens (723 K). An undoped PbS specimen produced at a sintering temperature of 723 K was too porous for comparison to the nearly dense Nadoped specimens, with a density of 6.27 g/cm3, and a P of 0.18, far exceeding the maximum of P < 0.05 for all other specimens in this study. Therefore, the porous undoped PbS specimen was excluded from the elasticity and hardness testing. 8.3.2 Elasticity results For the PbS-based specimens, the elastic moduli, E and G, were not sensitive to the addition of CdS or ZnS, with a mean and standard deviation of E of 63.38 ± 2.25 GPa (Figure 8.8a) and G was 24.69 ± 0.93 GPa (Figure 8.8c). The undoped PbS specimen had a much lower porosity, P = 0.008, than all the othere Na-doped PbS-based specimens, P = 0.031 to 0.050 (Table 8.1), likely due to the higher sintering temperature of the PbS specimen (see section 8.2.1 for sintering procedures). This difference in P likely explains the higher E and G of the undoped PbS specimen relative to the other PbS-based specimens (Figure 8.8a), and not a sensitivity to Na doping, as discussed in section 8.4.2. In contrast, the moduli of the PbSe-based specimens are sensitive to Na doping, but not CdS or ZnS additions. Both the E and G of the PbSe specimens decreased with the addition of Na dopant, from 64.29 ± 0.16 GPa and 25.29 ± 0.02 GPa for pure PbSe, to 60.33 ± 0.26 GPa and 23.66 ± 0.04 GPa for Na-doped PbSe (Figure 8.8b and 8.8d). No change in modulus was observed in the Na-doped PbSe specimens as a function of increasing additions of 1 at% to 4 at% CdS or ZnS except for a slight decrease in G and a corresponding increase in ν. The mean of E for the PbSe specimens with either CdS or 245 (a) 70 60 E with CdS E with ZnS E undoped E Na-doped 50 0 1 2 3 PbSe E (GPa) PbS E (GPa) 70 (b) 60 E with CdS E with ZnS E undoped E Na-doped 50 4 0 G with CdS G with ZnS G undoped G Na-doped 26 24 28 PbSe G (GPa) PbS G (GPa) 28 22 3 4 (d) G with CdS G with ZnS G undoped G Na-doped 26 24 22 0 1 2 3 4 0 % CdS/ZnS 0.40 0.35 0.30 0.25 0 1 2 3 1 2 3 4 % CdS/ZnS Poisson's ratio with CdS Poisson's ratio with ZnS Poisson's ratio undoped Poisson's ratio Na-doped (f) Poisson's ratio with CdS Poisson's ratio with ZnS Poisson's ratio undoped Poisson's ratio Na-doped 0.40 PbSe Poisson's ratio (e) PbS Poisson's ratio 2 % CdS/ZnS % CdS/ZnS (c) 1 0.35 0.30 0.25 0 4 1 2 3 4 % CdS/ZnS % CdS/ZnS Figure 8.8. The E and G of the undoped PbS and PbSe were higher than the Na-doped PbS and PbSe. The E and G of undoped PbS are likely higher due to reduced porosity (pure PbS P = 0.008, all other PbS P = 0.031 to 0.050), while the E and G of undoped PbSe are likely higher due to doping effects. Only small changes in E and G were noted in Na-doped PbS and PbSe with addition of up to 4% CdS or ZnS, with. Only small variability in the ν noted in PbS with addition of CdS or ZnS. The ν of PbSe increased with the addition of CdS or ZnS. Solid line is average and dotted line is standard deviation of the specimens with 1% to 4% CdS or ZnS addition. 246 (b) (a) 1.0 H with CdS H with ZnS H undoped H Na-doped 0.5 PbSe H (GPa) PbS H (GPa) 1.0 0.0 0.0 0 1 2 3 H with CdS H with ZnS H undoped H Na-doped 0.5 0 4 1 2 3 4 % CdS/ZnS % CdS/ZnS Figure 8.9. The hardness of both PbS and PbSe increased with the addition of Na dopant. No measurable change in hardness was noted in PbS with addition of up to 4% CdS or ZnS. Hardness of PbSe increased from 0.6 GPa to 0.8 GPa with the addition of either CdS or ZnS. 247 ZnS was 61.03 ± 0.79 and the mean of G was 23.35 ± 0.35 GPa) (Figure 8.8b, 8.8d, 8.8f). (Note: each of the specimens with additions of CdS or ZnS were Na-doped.) 8.3.3 Hardness results The hardness of PbS and PbSe increased with 2.5 at% or 2.0 at% Na-dopant additions, respectively, but were largely unaffected by the addition of 0 to 4 at% CdS or ZnS (Figure 8.9). The hardness of the pure PbS was 0.72 ± 0.10, and the average of the Na-doped PbS with additions of 0 to 4 at% CdS or ZnS was 1.12 ± 0.05 (Figure 8.9a). Similarly, the hardness of the pure PbSe was 0.44 ± 0.02 GPa, but increased to 0.62 ± 0.02 GPa when doped with 2.0 at% Na, and averaging 0.74 ± 0.08 GPa for specimens with 0 to 4 at% CdS or ZnS (Figure 8.9b). 8.4 Discussion 8.4.1 Microstructural analysis The observation of ZnS precipitates of up to 10 µm to 15 µm in diameter (Tables 8.1 and 8.2), as compared to CdS precipitates of up to 1.5 µm to 4 µm in both the PbS matrix (Table 8.1, Figure 8.4) and the PbSe matrix (Table 8.2, Figure 8.4), may be related to the much lower solubility limit of ZnS in the matrix [Zhao 2013]. More specifically for PbSe, “no solid solubility of Zn in PbSe was observed” [Zhao 2013]. Zhao et al. expected a continuous change in band gap up to the solubility limit for additions of CdS or ZnS in PbSe [Zhao 2013]. The band gap increased with up to 2% addition of CdS in PbSe, but with ZnS in PbSe, “no band gap changes are observed, suggesting a much lower solubility limit” of ZnS in PbSe [Zhao 2013]. Transmission electron microscopy observations of nanoprecipitates of CdS and ZnS in PbS and PbSe matrices have shown endotaxial crystallographic alignment with 248 the matrix [Zhao 2013; Zhao 2012a]. The morphology of the observed micron-scale precipitates with rod or plate geometry and sharp facets (Figure 8.4), particularly in the samples with CdS addition (Figure 8.5), are consistent with the observations of nanoprecipitate morphology. The change in fracture mode observed between specimens with and without CdS or ZnS addition (Tables 8.1 and 8.2) is often associated with a change in fracture toughness. A change from intergranular to transgranular fracture has been associated with either an increase [Mukhopadhyay 2010; Kawabata 1977; Karakasidis 2011] or decrease [Yamada 2010; Jang 2010]. The fracture toughness of the material was not determined in this study, but may be an area for future studies of PbS and PbSe with additions of CdS or ZnS. For PbS-based specimens sintered at 723 K, the P for the undoped material was 0.18, while the P of each of the Na-doped specimens was less than 0.05. A nearly dense (P = 0.01) undoped PbS specimen was sintered for this study, but only by increasing the sintering temperature to 873 K. The enhanced densification of the Na-doped PbS specimens may indicate the Na dopant acts as a sintering aid. A previous study on a different TE material, Ba0.3Co4Sb12, showed the addition of metal particles, specifically 0.5 wt% Ag nanoparticles, may act as sintering aid [Schmidt 2014]. In Ba0.3Co4Sb12, the Ag had a eutectic or peritectic reaction with Sb in the Ba0.3Co4Sb12 [Schmidt 2014] and a similar eutectic reaction may explain the enhanced densification in the Na-doped PbS specimens. The Na-S system has eutectics at 513 K and 61.5 at% Na, 522 K and 70 at% Na, and at 525 K and 72.5 at% Na [Sangster 1997]. All three of these eutectic points are well 249 below the sintering temperature used for the Na-doped PbS samples, 723 K. In the closely related PbTe material system, the Na in Na-doped PbTe partially segregates in a layer less than 10 nm thick at the grain boundaries [Biswas 2012]. A similar Na-rich layer may also form during sintering in the PbS-based material in this study, and this Narich layer may act as a liquid-phase sintering aid for the PbS-based system that accounts for the enhanced sintering. The sintering behavior of Na-doped PbS has not been explored and may be a topic for future work. 8.4.2 Elasticity analysis For the PbS-based specimens, the undoped PbS specimen exhibited higher E and G relative to the other PbS-based specimens (Figure 8.8a). The difference in elastic moduli between PbS-based specimens likely a function of porosity, as the porosity of the undoped PbS specimen is P = 0.008, much lower than all the other Na-doped PbS-based specimens, P = 0.031 to 0.050 (Table 8.1), and not because of a sensitivity to the Na doping. In general, the elastic modulus of TE materials and other brittle materials is a linear function of porosity [Schmidt 2013a; Ni 2009b; Rice 1998; Ni 2011], E  E0 1  bE P  (4) where E0 is the Young’s modulus at P = 0, and bE is a measure of the decrease in E with increasing P. For two previously studied TE materials, YbAl3 and LAST (lead– antimony–silver–tellurium), bE is 2.34 [Schmidt 2013a] and 3.6 [Ni 2009b]. Based on these values of bE, the difference between the undoped PbS and the Na-doped specimens with 0 to 4 at% addition of CdS or ZnS is expected to be between 4 GPa and 11 GPa, and consistent with the measured difference of approximately 6 GPa (Figure 8.8a). 250 The elastic moduli for the doped TE materials with nanoprecipitates in this study have not been reported, however the moduli of the undoped PbS [Dalven 1969; Bhagavantam 1951] and PbSe [Lippmann 1971; Dalven 1969; Hellwege 1979] are known for single crystal specimens. The aggregate average of the Hashin and Shtrikman bounds, , [Simmons 1971] allow a direct comparison of the single crystal specimen moduli to the moduli of a dense polycrystalline specimen (Table 8.4). Comparing this study to the averages, the polycrystalline PbS specimen in this study (which has a volume fraction porosity, P of 0.008) were lower than the single crystal values by 12% to 18%, with the measured E of 68.95 ± 0.10 GPa in this study, and the single crystal E of 78.05 GPa [Dalven 1969] and 83.61 GPa [Bhagavantam 1951] (Table 8.4). The elastic moduli of PbSe in this study (P = 0.016), with E of 64.29 ± 0.16 GPa, generally agrees with previous studies, with E between 61.68 GPa and 70.69 GPa [Lippmann 1971; Dalven 1969; Hellwege 1979] (Table 8.4). The reader may ask, “why are the moduli of PbS lower than those reported in literature?” The two most likely causes would be (i) porosity and (ii) doping or composition differences. Porosity is not sufficient by itself to explain the difference between the literature values of E and the measured E in this study. The contribution from porosity may be estimated using equation 4 and applying the measured bE from two previously studied 251 Table 8.4. The Young’s modulus, E, the shear modulus, G, and the Poisson’s ratio measured in this study for the polycrystalline undoped PbSe and undoped PbS specimens compared with the aggregate average values of E and G computed from the Hashin and Shtrikman bounds, [Simmons 1971] for single crystal elasticity from the literature [Dalven 1969, Bhagavantam 1951, Lippmann 1971, Hellwege 1979]. Material Bulk material form PbS Polycrystalline PbS Single crystal E (GPa) 68.95 78.05 G (GPa) 27.16 29.98 Poisson’s ratio 0.269 0.302 PbS Single crystal 83.61 32.53 0.277 PbSe PbSe PbSe Polycrystalline Single crystal Single crystal 64.29 66.83 70.69 25.29 23.76 24.11 0.271 0.298 0.466 PbSe Single crystal 61.68 NA: Not available from literature 23.76 0.294 252 Density (g/cm3) 7.53 7.5 (P=0.01) NA Reference This study [Dalven 1969] [Bhagavantam 1951] This study [Lippmann 1971] [Dalven 1969] 8.13 8.26 8.16 (P=0.01) NA [Hellwege 1979] TE materials, YbAl3 (bE = 2.34) [Schmidt 2013a] and LAST (bE = 3.6) [Ni 2009b]. Based on these values of bE and the observed porosity in the PbS specimen (P = 0.046), the E0 of PbS in this study would likely be in the range of 70.3 GPa to 71.0 GPa, which is still 12% to 15% less than the average of literature values (Table 8.4). After accounting for porosity, the remaining differences between the moduli of the polycrystalline PbS specimen in this study and the literature may be related to composition or doping and vacancies. To examine this possibility, a cut surface of the PbS specimen was examined by EDS to determine an approximate composition (Table 8.3). The EDS surface composition of a polished surface of the specimen was PbS0.98, averaged over 5 separate sites, where the area of each site was approximately 0.05 mm2. The surface composition of the PbS specimen is thus consistent with sulfur vacancies (Table 8.3). In this study, the PbS was densified by PECS in an argon atmosphere using a starting powder milled from an ingot. During the initial stage of sintering before densification occurs, the surfaces of the PbS powder particles are exposed to the sintering atmosphere at elevated temperature, 873 K, providing a pathway for S loss. Prior to sintering, the powder was single phase in XRD, with a lattice parameter of 5.932 Å [Zhao 2012a]. After sintering, the specimen was single phase in XRD, but with a lattice parameter of 5.941 Å for the sintered specimen (Figure 8.7). The increase in lattice parameter is consistent with a change in vacancy concentration. The observed reduction in the elastic moduli by roughly 12% to 15% in this study may be caused by sulfur vacancies introduced during sintering, as indicated by EDS analysis. A change in elastic moduli by roughly 12% to 15% due to vacancies is not unique to PbS. Another thermoelectric, SnTe1+x, exhibits a decrease of E by 12% with 253 increasing Te content from x = 0 to 0.016 and a subsequent rise in the number of cation vacancies [Schmidt 2013b]. In addition, the elastic modulus of a set of nitrides decreases with increasing N vacancies. The E of TiNx coatings decreases by 23% [Shin 2003] or 41% [Portnoi 1968] for a change in x from 1 to 0.67, decreases ~20% [Jiang 1991] or 50% [Portnoi 1968] for x from 1 to 0.5, and by 5% for a small change of x from 0.98 to 0.94 [Portnoi 1968]. Similarly, the E of ZrNx decreased 12% for a change in x from 1 to 0.85 [Portnoi 1968]. For CeO2-x-based compounds, increasing O vacancies (x from approximately 0 to approximately 0.11, at O2 partial pressure of 0.22 atm and 4.50 × 10-22 atm) reduced the E of ceria by 11% and gadolinium doped ceria by 6% [Wang 2007]. 8.4.3 Hardness analysis Hardness values for single crystal PbS [Bloem 1955] and polycrystalline specimens of PbS and PbSe [Darrow 1969] are available in the literature (Table 8.5), but no hardness data is available in the literature for PbS and PbSe specimens optimized as TE materials by the inclusion of Na dopant and CdS or ZnS nano/micro precipitates. The hardness of PbS specimens included in this study was not observed to be a function of addition of precipitates or the dopant (Table 8.5, Figure 8.9a). The hardness of PbS is higher than reported in the literature [Bloem 1955; Darrow 1969], but hardness is known to be a function of grain size, and would be expected to be higher for the powder processed specimens in this study than for single crystal specimens of the same composition [Rice 1994]. 254 Table 8.5. In general, hardness, H, is grain size dependent, and the single crystal hardness results from Bloem and Kr ger [Bloem 1955] are expected to have a lower hardness than the other, polycrystalline specimens in this table. The grain size for the PbS and PbSe was not listed by Darrow [Darrow 1969]. Average grain sizes for all specimens in this study are between 1.8 µm and 18.7 µm, listed in Tables 8.1 and 8.2. Material Bulk material form PbS Polycrystalline PbS Single crystal PbS PbS PbSe Na-doped PbSe PbSe Indentation Reference Load (N) 1.96 This study 0.98 [Bloem 1955] Polycrystalline Polycrystalline Polycrystalline Polycrystalline H (GPa) 1.09 ± 0.14 0.4 to 0.7 (dependent on vacancy concentration) 0.72 to 0.92 0.73 0.44 ± 0.02 0.62 ± 0.02 1.41 3.92 1.96 1.96 [Darrow 1969] [Darrow 1969] This study This study Polycrystalline 0.56 to 0.57 1.41 [Darrow 1969] 255 Although the hardness was not observed to be a function of Na doping in this study, the hardness of PbS is a function of the vacancy concentration [Bloem 1955]. On single crystal specimens of PbS, Bloem and Kroger reported a minimum hardness of 0.4 GPa is reported when there is a minimum of vacancies [Bloem 1955]. In this study, the PbS was S-deficient (Table 8.3), implying S vacancies were present, regardless of the Na doping and regardless of whether or not CdS or ZnS was added to the matrix. The difference between the constant H in this study and the minimum reported by Bloem and Kroger [Bloem 1955] is likely because there are likely a large number of S vacancies (Table 8.3) in each of the PbS specimens in this study. For future work with different processing conditions or Na doping concentrations, the vacancy concentration may approach a value where a minimum in H is possible, and the value of H should be reexamined. For PbSe, there may be a similar relationship between vacancies and hardness as was observed by Bloem and Kroger in PbS [Bloem 1955]. Without CdS or ZnS additions, the hardness increased from 0.44 ± 0.02 GPa to 0.62 ± 0.02 GPa with the addition of 2.0 at% Na dopant (Table 8.5, Figure 8.9b). After the addition of Na, the hardness was only a weak function, if any, of CdS or ZnS addition, averaging 0.75 ± 0.07 GPa for all the specimens with CdS or ZnS addition (Figure 8.9b). Fracture toughness was not measured in this study. For each of the specimens included in this study, Vickers indentation produced an impression with a set of typically 8 to 15 or more cracks surrounding and extending out from the sides of the indentation impression, rather than a set of radial or Palmquist cracks that extend from the corners of the indentation impression. The lack of well-defined radial cracks 256 precluded making an estimate of the fracture toughness for the materials in this study. These cracks may develop as a result of many mechanisms, such as the environment, microstructural toughening, surface stress, or preferred crystallographic cleavage planes [Cook 2003]. Future studies may examine fracture toughness by a method other than Vickers indentation. 8.5 Conclusions For PbS + x at%(CdS, ZnS), neither the elastic moduli, with E averaging 63.38 ± 2.25 GPa, nor the H, averaging 1.12 ± 0.05 GPa, are a function of addition of 2.5 at% Na dopant and 0 to 4 at % CdS or ZnS, but the elastic moduli of the sintered pure PbS is lower than previous reports. The lower elastic modulus is partially accounted for by porosity, but may also be related to S vacancies. For PbSe + x at%(CdS, ZnS), the elastic moduli and hardness of PbSe are a function of 2.0 at% Na dopant, but not a function of addition of 0 to 4 at% CdS or ZnS. Without Na dopant, the E was 64.29 ± 0.16 GPa and H was 0.44 ± 0.02 GPa. With the Na dopant, the E averaged 61.03 ± 0.79 and the H averaged 0.75 ± 0.07 GPa. The elastic moduli are required material properties to model stresses of a TE material in a module or device, and stress modeling is necessary for designing or fabrication of a mechanically reliable device. Hardness is related to wear characteristics for a material. As the transport properties of a TE material are optimized, the mechanical properties may change, potentially requiring a redesign of a module or device. The continuing work on optimization of PbS- [Zhao 2012a] or PbSe-based [Zhao 2013] TE materials carries implications on changing the mechanical properties of these materials required for designing and modeling a TE module or device. The effects on 257 mechanical properties from two aspects of optimization were examined in this study, (i) addition of a precipitate phase, namely CdS or ZnS, and (ii) changes based on Na doping or S deficiency. Optimizing either PbS or PbSe by addition of up to 4 at% CdS or ZnS may be performed without concern for changing elastic moduli or H, nor the mechanical design of TE devices reliant upon the elastic moduli or H, although a change in fracture mode may indicate a change in fracture toughness. The addition of 1 at% to 4 at% CdS or ZnS to either a PbS or PbSe matrix changes the fracture mode of the specimen, from primarily intergranular fracture to primarily transgranular fracture. A change in mode mixity is often associated with a change in fracture toughness. Fracture toughness could not be determined by Vickers indentation in this study, but may be measured by an alternative method in future work. The CdS and ZnS additions were observed as precipitates, with the range of precipitate diameter generally not a function of concentration or if the matrix was PbS or PbSe. The maximum diameter of CdS precipitates is up to 4 µm, and the maximum size of ZnS precipitates up to 15 µm. ACKNOWLEDGEMENTS The authors acknowledge the financial support of the Department of Energy, Revolutionary Materials for Solid State Energy Conversion Center, an Energy Frontiers Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under award number DE-SC0001054. 258 REFERENCES 259 REFERENCES [Alleno, Chen, Chubilleau, Lenoir, Rouleau, Trichet, Villeroy 2009] Alleno, E., Chen, L., Chubilleau, C., Lenoir, B., Rouleau, O., Trichet, M.F., Villeroy, B. Thermal Conductivity Reduction in CoSb3–CeO2 Nanocomposites. Journal of Electronic Materials 39 (2009) 1966–1970. [Androulakis, Lin, Kong, Uher, Wu, Hogan, Cook, Caillat, Paraskevopoulos, Kanatzidis 2007] Androulakis, J., Lin, C.-H., Kong, H.-J., Uher, C., Wu, C.-I., Hogan, T., Cook, B.A., Caillat, T., Paraskevopoulos, K.M., Kanatzidis, M.G. Spinodal decomposition and nucleation and growth as a means to bulk nanostructured thermoelectrics: enhanced performance in Pb(1-x)Sn(x)Te-PbS. Journal of the American Chemical Society 129 (2007) 9780–8. [Beattie 1969] Beattie, A.G. Temperature Dependence of the Elastic Constants of Tin Telluride. Journal of Applied Physics 40 (1969) 4818. [Bhagavantam, Rao 1951] Bhagavantam, S., Rao, T. Elastic constants of galena. Nature 168 (1951) 42. [Bhattacharya, Hermann, Keppens, Snyder 2006] Bhattacharya, S., Hermann, R.P., Keppens, V., Snyder, G.J. Effect of disorder on the thermal transport and elastic properties in thermoelectric Zn_4Sb_3. Physical Review B 74 (2006) 1–5. [Biswas, He, Blum, Wu, Hogan, Seidman, Dravid, Kanatzidis 2012] Biswas, K., He, J., Blum, I.D., Wu, C.-I., Hogan, T.P., Seidman, D.N., Dravid, V.P., Kanatzidis, M.G. High-performance bulk thermoelectrics with all-scale hierarchical architectures. Nature 489 (2012) 414–8. [Bloem, Kröger 1955] Bloem, J., Kröger, F. A Relation between Hardness and Stoichiometry in Lead Sulphide Single Crystals. Nature 175 (1955) 861. [Cline, Dunegan, Henderson 1967] Cline, C.F., Dunegan, H.L., Henderson, G.W. Elastic Constants of Hexagonal BeO, ZnS, and CdSe. Journal of Applied Physics 38 (1967) 1967. [Cook 2003] Cook, R.F. Microstructural Control of Indentation Crack Extension under Externally Applied Stress. Fracture mechanics of ceramics 14 (2003) 57–67. [Dalven 1969] Dalven, R. A review of the semiconductor properties of PbTe, PbSe, PbS and PbO. Infrared physics 9 (1969) 141–184. 260 [Darrow, White, Roy 1969] Darrow, M., White, W., Roy, R. Micro-indentation hardness variation as a function of composition for polycrystalline solutions in the systems PbS/PbTe, PbSe/PbTe, and PbS/PbSe. Journal of Materials Science 4 (1969) 313– 319. [Fan, Case, Lu, Morelli 2013] Fan, X., Case, E.D., Lu, X., Morelli, D.T. Room temperature mechanical properties of natural-mineral-based thermoelectrics. Journal of Materials Science 48 (2013) 7540–7550. [Gelbstein, Gotesman, Lishzinker, Dashevsky, Dariel 2008] Gelbstein, Y., Gotesman, G., Lishzinker, Y., Dashevsky, Z., Dariel, M.P. Mechanical properties of PbTe-based thermoelectric semiconductors. Scripta Materialia 58 (2008) 251–254. [Haxel, Hedrick, Orris 2002] Haxel, G.B., Hedrick, J.B., Orris, G.J. Rare Earth Elements — Critical Resources for High Technology. USGS Fact Sheet fs087-02, (2002) . http://pubs.usgs.gov/fs/2002/fs087-02. [Hellwege 1979] Landold-Bornstein numerical data and functional relationships in science and technology, new series, Volume 11. Springer-Verlag, Berlin, 1979. [Heremans, Jovovic, Toberer, Saramat, Kurosaki, Charoenphakdee, Yamanaka, Snyder 2008] Heremans, J.P., Jovovic, V., Toberer, E.S., Saramat, A., Kurosaki, K., Charoenphakdee, A., Yamanaka, S., Snyder, G.J. Enhancement of thermoelectric efficiency in PbTe by distortion of the electronic density of states. Science (New York, N.Y.) 321 (2008) 554–7. [Houston, Strakna, Belson 1968] Houston, B., Strakna, R.E., Belson, H.S. Elastic Constants, Thermal Expansion, and Debye Temperature of Lead Telluride. Journal of Applied Physics 39 (1968) 3913. [Hsu, Loo, Guo, Chen, Dyck, Uher, Hogan, Polychroniadis, Kanatzidis 2004] Hsu, K.F., Loo, S., Guo, F., Chen, W., Dyck, J.S., Uher, C., Hogan, T., Polychroniadis, E.K., Kanatzidis, M.G. Cubic AgPb(m)SbTe(2+m): bulk thermoelectric materials with high figure of merit. Science (New York, N.Y.) 303 (2004) 818–21. [Jang, Kim, Han, Seo, Hong, Woo, Sakka 2010] Jang, B.-K., Kim, S.-Y., Han, I.-S., Seo, D.-W., Hong, K.-S., Woo, S.-K., Sakka, Y. Influence of uni and bi-modal SiC composition on mechanical properties and microstructure of reaction-bonded SiC ceramics. Journal of the Ceramic Society of Japan 118 (2010) 1028–1031. [Ji, He, Alboni, Su, Gothard, Zhang, Tritt, Kolis 2007] Ji, X., He, J., Alboni, P., Su, Z., Gothard, N., Zhang, B., Tritt, T.M., Kolis, J.W. Thermal conductivity of CoSb3 nano-composites grown via a novel solvothermal nano-plating technique. physica status solidi RRL – Rapid Research Letters 1 (2007) 229–231. 261 [Jiang, Wang, Schmidt, Dunlop, Haupt, Gissler 1991] Jiang, X., Wang, M., Schmidt, K., Dunlop, E., Haupt, J., Gissler, W. Elastic constants and hardness of ion-beamsputtered TiNx films measured by Brillouin scattering and depth-sensing indentation. Journal of Applied Physics 69 (1991) 3053–3057. [Karakasidis, Charitidis 2011] Karakasidis, T.E., Charitidis, C. a. Influence of nanoinclusions’ grain boundaries on crack propagation modes in materials. Materials Science and Engineering: B 176 (2011) 490–493. [Kawabata, Izumi 1977] Kawabata, T., Izumi, O. The relationship between fracture toughness and transgranular fracture in an Al-6.0% Zn-2.5% Mg alloy. Acta Metallurgica 25 (1977) 505–512. [Kvetková, Duszová, Kašiarová, Dorčáková, Dusza, Balázsi 2013] Kvetková, L., Duszová, A., Kašiarová, M., Dorčáková, F., Dusza, J., Balázsi, C. Influence of processing on fracture toughness of Si3N4+graphene platelet composites. Journal of the European Ceramic Society 33 (2013) 2299–2304. [Lippmann, Kastner, Wanninger 1971] Lippmann, G., Kastner, P., Wanninger, W. Elastic constants of PbSe. physica status solidi (a) 159 (1971) 159–161. [Mi, Zhao, Zhu, Tu 2008] Mi, J.L., Zhao, X.B., Zhu, T.J., Tu, J.P. Thermoelectric properties of Yb 0.15 Co 4 Sb 12 based nanocomposites with CoSb 3 nanoinclusion. Journal of Physics D: Applied Physics 41 (2008) 205403. [Migliori, Sarrao 1997] Migliori, A., Sarrao, J.L. Resonant Ultrasound Spectroscopy: Applications to Physics, Materials Measurements, and Nondestructive Evaluation. Wiley-Vch, New York, 1997. [Milekhine, Onsøien, Solberg, Skaland 2002] Milekhine, V., Onsøien, M.I., Solberg, J.K., Skaland, T. Mechanical properties of FeSi (e), FeSi2 (za) and Mg2Si. Intermetallics 10 (2002) 743–750. [Möchel, Sergueev, Wille, Voigt, Prager, Stone, Sales, Guguchia, Shengelaya, Keppens, Hermann 2011] Möchel, a., Sergueev, I., Wille, H.-C., Voigt, J., Prager, M., Stone, M.B., Sales, B.C., Guguchia, Z., Shengelaya, a., Keppens, V., Hermann, R.P. Lattice dynamics and anomalous softening in the YbFe_{4}Sb_{12} skutterudite. Physical Review B 84 (2011) 184306. [Morrison, Case, Ren, Baumann, Kleinow, Ni, Hogan, D’Angelo, Matchanov, Hendricks, Karri, Cauchy, Barnard, Kanatzidis 2012] Morrison, a. Q., Case, E.D., Ren, F., Baumann, a. J., Kleinow, D.C., Ni, J.E., Hogan, T.P., D’Angelo, J., Matchanov, N. a., Hendricks, T.J., Karri, N.K., Cauchy, C., Barnard, J., Kanatzidis, M.G. Elastic modulus, biaxial fracture strength, electrical and thermal transport properties of 262 thermally fatigued hot pressed LAST and LASTT thermoelectric materials. Materials Chemistry and Physics 134 (2012) 973–987. [Mukhopadhyay, Chakravarty, Basu 2010] Mukhopadhyay, A., Chakravarty, D., Basu, B. Spark Plasma-Sintered WC-ZrO 2 -Co Nanocomposites with High Fracture Toughness and Strength. Journal of the American Ceramic Society 93 (2010) 1754– 1763. [Ni, Ren, Case, Timm 2009a] Ni, J., Ren, F., Case, E., Timm, E. Porosity dependence of elastic moduli in LAST (lead–antimony–silver–tellurium) thermoelectric materials. Materials Chemistry and Physics(2009) . [Ni, Case, Khabir, Stewart, Wu, Hogan, Timm, Girard, Kanatzidis 2010] Ni, J.E., Case, E.D., Khabir, K.N., Stewart, R.C., Wu, C.-I., Hogan, T.P., Timm, E.J., Girard, S.N., Kanatzidis, M.G. Room temperature Young’s modulus, shear modulus, Poisson's ratio and hardness of PbTe–PbS thermoelectric materials. Materials Science and Engineering: B 170 (2010) 58–66. [Ni, Case, Schmidt, Wu, Hogan, Trejo, Lara-Curzio, Kanatzidis 2013] Ni, J.E., Case, E.D., Schmidt, R.D., Wu, C.-I., Hogan, T.P., Trejo, R.M., Lara-Curzio, E., Kanatzidis, M.G. Fracture mode, microstructure and temperature-dependent elastic moduli for thermoelectric composites of PbTe–PbS with SiC nanoparticle additions. Philosophical Magazine 93 (2013) 4412–4439. [Ni, Case, Stewart, Wu, Hogan, Kanatzidis 2011] Ni, J.E., Case, E.D., Stewart, R., Wu, C.-I., Hogan, T.P., Kanatzidis, M.G. Bloating in (Pb0.95Sn0.05Te)0.92(PbS)0.080.055%PbI2 Thermoelectric Specimens as a Result of Processing Conditions. Journal of Electronic Materials 41 (2011) 1153–1158. [Ni, Ren, Case, Timm 2009b] Ni, J.E., Ren, F., Case, E.E.D., Timm, E.J.E. Porosity dependence of elastic moduli in LAST (lead–antimony–silver–tellurium) thermoelectric materials. Materials Chemistry and Physics 118 (2009) 459–466. [Portnoi, Mukaseev, Gribkov, Levinskii 1968] Portnoi, I., Mukaseev, A.A., Gribkov, V.N., Levinskii, Y. V. Modulus of normal elasticity of porosity-free titanium and zirconium nitrides. Soviet Powder Metallurgy and Metal Ceramics 7 (1968) 406– 408. [Ren, Case, Ni, Timm, Lara-Curzio, Trejo, Lin, Kanatzidis 2009] Ren, F., Case, E.D., Ni, J.E., Timm, E.J., Lara-Curzio, E., Trejo, R.M., Lin, C.-H., Kanatzidis, M.G. Temperature-dependent elastic moduli of lead telluride-based thermoelectric materials. Philosophical Magazine 89 (2009) 143–167. [Ren, Case, Timm, Lara-Curzio, Trejo 2010] Ren, F., Case, E.D., Timm, E.J., LaraCurzio, E., Trejo, R.M. Anomalous temperature-dependent Young’s modulus of a 263 cast LAST (Pb–Sb–Ag–Te) thermoelectric material. Acta Materialia 58 (2010) 31– 38. [Ren, Case, Timm, Schock 2007] Ren, F., Case, E.D., Timm, E.J., Schock, H.J. Young’s modulus as a function of composition for an n-type lead–antimony–silver–telluride (LAST) thermoelectric material. Philosophical Magazine 87 (2007) 4907–4934. [Ren, Hall, Ni, Case, Sootsman, Mercouri 2008] Ren, F., Hall, B.D., Ni, J.E., Case, E.D., Sootsman, J., Mercouri, G. Mechanical Characterization of PbTe-based Thermoelectric Materials. 1044 (2008) . [Rice 1998] Rice, R.W. Porosity of Ceramics. Marcel Dekker, New York, 1998. [Rice, Wu, Borchelt 1994] Rice, R.W., Wu, C.C., Borchelt, F. Hardness–Grain-Size Relations in Ceramics. Journal of the American Ceramic Society 77 (1994) 2539– 2553. [Rogl, Rogl 2011] Rogl, G., Rogl, P. Mechanical Properties of Skutterudites. Science of Advanced Materials 3 (2011) 517–538. [Salvador, Yang, Shi, Wang, Wereszczak, Kong, Uher 2009] Salvador, J.R., Yang, J., Shi, X., Wang, H., Wereszczak, a. a., Kong, H., Uher, C. Transport and mechanical properties of Yb-filled skutterudites. Philosophical Magazine 89 (2009) 1517–1534. [Sangster, Pelton 1997] Sangster, J., Pelton, A. The Na-S (sodium-sulfur) system. Journal of phase equilibria 18 (1997) 89–96. [Schmidt, Case, Giles, Ni, Hogan 2012a] Schmidt, R.D., Case, E.D., Giles, J., Ni, J.E., Hogan, T.P. Room-Temperature Mechanical Properties and Slow Crack Growth Behavior of Mg2Si Thermoelectric Materials. Journal of Electronic Materials 41 (2012) 1210–1216. [Schmidt, Case, Lehr, Morelli 2013a] Schmidt, R.D., Case, E.D., Lehr, G.J., Morelli, D.T. Room temperature mechanical properties of polycrystalline YbAl3, a promising low temperature thermoelectric material. Intermetallics 35 (2013) 15–24. [Schmidt, Case, Lobo, Thompson, Sakamoto, Zhou, Uher 2014] Schmidt, R.D., Case, E.D., Lobo, Z., Thompson, T.R., Sakamoto, J.S., Zhou, X.-Y., Uher, C. Influence of silver nanoparticle addition, porosity, and processing technique on the mechanical properties of Ba0.3Co4Sb12 skutterudites. Journal of Materials Science(2014) . [Schmidt, Case, Ni, Sakamoto, Trejo, Lara-Curzio 2012b] Schmidt, R.D., Case, E.D., Ni, J.E., Sakamoto, J.S., Trejo, R.M., Lara-Curzio, E. Temperature-dependent Young’s modulus, shear modulus and Poisson's ratio of p -type Ce 0.9 Fe 3.5 Co 0.5 Sb 12 264 and n -type Co 0.95 Pd 0.05 Te 0.05 Sb 3 skutterudite thermoelectric materials. Philosophical Magazine 92 (2012) 727–759. [Schmidt, Case, Ni, Trejo, Lara-Curzio, Korkosz, Kanatzidis 2013b] Schmidt, R.D., Case, E.D., Ni, J.E., Trejo, R.M., Lara-Curzio, E., Korkosz, R.J., Kanatzidis, M.G. High-temperature elastic moduli of thermoelectric SnTe1±x – y SiC nanoparticulate composites. Journal of Materials Science 48 (2013) 8244–8258. [Schmidt, Ni, Case, Sakamoto, Kleinow, Wing, Stewart, Timm 2010] Schmidt, R.D., Ni, J.E., Case, E.D., Sakamoto, J.S., Kleinow, D.C., Wing, B.L., Stewart, R.C., Timm, E.J. Room temperature Young’s modulus, shear modulus, and Poisson's ratio of Ce0.9Fe3.5Co0.5Sb12 and Co0.95Pd0.05Te0.05Sb3 skutterudite materials. Journal of Alloys and Compounds 504 (2010) 303–309. [Shin, Gall, Hellgren, Patscheider, Petrov, Greene 2003] Shin, C.-S., Gall, D., Hellgren, N., Patscheider, J., Petrov, I., Greene, J.E. Vacancy hardening in single-crystal TiN[sub x](001) layers. Journal of Applied Physics 93 (2003) 6025–6028. [Simmons, Wang 1971] Simmons, G., Wang, H. Single Crystal Elastic Constants and Calculated Aggregate Properties: A Handbook. The M.I.T. Press, Cambridge, Massachusetts, 1971. [Toprak, Stiewe, Platzek, Williams, Bertini, Muller, Gatti, Zhang, Rowe, Muhammed 2004] Toprak, M.S., Stiewe, C., Platzek, D., Williams, S., Bertini, L., Muller, E., Gatti, C., Zhang, Y., Rowe, M., Muhammed, M. The Impact of Nanostructuring on the Thermal Conductivity of Thermoelectric CoSb3. Advanced Functional Materials 14 (2004) 1189–1196. [Tritt 2011] Tritt, T.M. Thermoelectric Phenomena, Materials, and Applications. Annual Review of Materials Research 41 (2011) 433–448. [Wang, Duncan, Wachsman, Ebrahimi 2007] Wang, Y., Duncan, K., Wachsman, E., Ebrahimi, F. The effect of oxygen vacancy concentration on the elastic modulus of fluorite-structured oxides. Solid State Ionics 178 (2007) 53–58. [Whitten, Chung, Danielson 1965] Whitten, W., Chung, P., Danielson, G. Elastic constants and lattice vibration frequencies of Mg 2 Si. Journal of Physics and Chemistry of Solids 26 (1965) 49–56. [Yamada, Sekine, Kumazawa, Tanabe 2010] Yamada, M., Sekine, K., Kumazawa, T., Tanabe, Y. Relationship between the cone crack and fracture mode in ceramics under high-velocity-projectile impact. 96 (2010) 903–908. [Zhang, Rogl, Grytsiv, Puchegger, Koppensteiner, Spieckermann, Kabelka, Reinecker, Rogl, Schranz, Zehetbauer, Carpenter 2010] Zhang, L., Rogl, G., Grytsiv, A., 265 Puchegger, S., Koppensteiner, J., Spieckermann, F., Kabelka, H., Reinecker, M., Rogl, P., Schranz, W., Zehetbauer, M., Carpenter, M.A. Mechanical properties of filled antimonide skutterudites. Materials Science and Engineering: B 170 (2010) 26–31. [Zhao, Hao, Lo, Wu, Zhou, Lee, Li, Biswas, Hogan, Uher, Wolverton, Dravid, Kanatzidis 2013] Zhao, L.-D., Hao, S., Lo, S.-H., Wu, C.-I., Zhou, X., Lee, Y., Li, H., Biswas, K., Hogan, T.P., Uher, C., Wolverton, C., Dravid, V.P., Kanatzidis, M.G. High thermoelectric performance via hierarchical compositionally alloyed nanostructures. Journal of the American Chemical Society 135 (2013) 7364–70. [Zhao, He, Hao, Wu, Hogan, Wolverton, Dravid, Kanatzidis 2012a] Zhao, L.-D., He, J., Hao, S., Wu, C.-I., Hogan, T.P., Wolverton, C., Dravid, V.P., Kanatzidis, M.G. Raising the thermoelectric performance of p-type PbS with endotaxial nanostructuring and valence-band offset engineering using CdS and ZnS. Journal of the American Chemical Society 134 (2012) 16327–36. [Zhao, He, Wu, Hogan, Zhou, Uher, Dravid, Kanatzidis 2012b] Zhao, L.-D., He, J., Wu, C.-I., Hogan, T.P., Zhou, X., Uher, C., Dravid, V.P., Kanatzidis, M.G. Thermoelectrics with earth abundant elements: high performance p-type PbS nanostructured with SrS and CaS. Journal of the American Chemical Society 134 (2012) 7902–12. [Zhao, Lo, He, Li, Biswas, Androulakis, Wu, Hogan, Chung, Dravid, Kanatzidis 2011] Zhao, L.-D., Lo, S.-H., He, J., Li, H., Biswas, K., Androulakis, J., Wu, C.-I., Hogan, T.P., Chung, D.-Y., Dravid, V.P., Kanatzidis, M.G. High performance thermoelectrics from earth-abundant materials: enhanced figure of merit in PbS by second phase nanostructures. Journal of the American Chemical Society 133 (2011) 20476–87. [Zhao, Zhang, Li, Zhou, Liu, Liu 2008] Zhao, L.-D., Zhang, B.-P., Li, J.-F., Zhou, M., Liu, W.-S., Liu, J. Thermoelectric and mechanical properties of nano-SiC-dispersed Bi2Te3 fabricated by mechanical alloying and spark plasma sintering. Journal of Alloys and Compounds 455 (2008) 259–264. [Zhou, Li, Kita 2008] Zhou, M., Li, J.-F., Kita, T. Nanostructured AgPb(m)SbTe(m+2) system bulk materials with enhanced thermoelectric performance. Journal of the American Chemical Society 130 (2008) 4527–32. [Zhu, Lee, Lan, Wang, Joshi, Wang, Yang, Vashaee, Guilbert, Pillitteri, Dresselhaus, Chen, Ren 2009] Zhu, G., Lee, H., Lan, Y., Wang, X., Joshi, G., Wang, D., Yang, J., Vashaee, D., Guilbert, H., Pillitteri, a., Dresselhaus, M., Chen, G., Ren, Z. Increased Phonon Scattering by Nanograins and Point Defects in Nanostructured Silicon with a Low Concentration of Germanium. Physical Review Letters 102 (2009) 196803. 266 9 Summary and Conclusions 9.1 Porosity dependence and processing method independence of elastic moduli From the results of Chapter 3 on properties of YbAl3 and Chapter 6 on Ba0.3Co4Sb12 skutterudites, we conclude that the Young’s modulus, E, shear modulus, G, and bulk modulus, B, of bulk porous YbAl3 and Ba0.3Co4Sb12 are a function primarily of the volume fraction of natural porosity, P. In this work, natural porosity refers to porosity that occurs when powder processing and sintering a brittle material, as opposed to introduced porosity such as that formed when a second material is mixed and then burned out during sintering or when pores are formed via bloating. Two powder processing and sintering methods, (i) grinding and sieving and hot press (HP) sintering, and (ii) vibratory milling and pulsed electric current sintering (PECS), were used to produce sets of specimens with a range of porosity for each of YbAl3 and Ba0.3Co4Sb12. The powder processing method, sintering method, or different natural porosity size and distribution are not a significant factor in determining the mechanical properties of E, G, and H for YbAl3 and Ba0.3Co4Sb12. The properties may be accurately estimated with only a knowledge of P. The elastic moduli, E, G, and B, of many brittle materials have been shown follow the empirical relationship [Rice 1998] A = A0(exp-bAP) (1) where A represents E, G, or B, A0 represents E0, G0 and B0 , the P = 0 intercepts, and bA is a unitless, material-dependent parameter. If bAP is small, equation (1) may be linearized [Ni 2009b] by using the first two terms of the Taylor series expansion of equation (1). 267 A = A0(1 - bAP) (2) For both YbAl3 and Ba0.3Co4Sb12, the linear form has been used. Rice [Rice 1998] has shown that the porosity dependence, bA, of the elastic moduli for a variety of brittle materials typically lies between 2 and 6. Prior to this work, the bE and bG values for only one TE material was in the literature, LAST (lead– antimony–silver–tellurium) [Ni 2009a]. 9.1.1 Grain size and processing method independence of elastic moduli Processing conditions may lead to a change in grain size (GS), but a change in GS does not change the elastic moduli for specimens with GS greater than approximately 20 nm [Kim 1999]. This was consistent with the results for the Ba0.3Co4Sb12 specimens, where processing method changed the Ba0.3Co4Sb12 GS from bimodal, with most grains >1 µm in diameter for the HP, to unimodal with an average GS of 0.14 – 0.17 µm. The elastic moduli results for the different GS Ba0.3Co4Sb12 specimens in section 6 further support the result that only the P dependence is required to predict the E, G, or B of a TE material. 9.2 Sintering aids and TE materials Sintering aids have generally not been considered for TE materials, despite the numerous benefits, such as reduced sintering time, temperature and cost [German 1985]. A new possible sintering aid was determined for Sb-based skutterudites and other Sbcontaining TE materials, as well as a method to identify possible sintering aids for other TE materials (chapter 6). In examining the increased density of Ba0.3Co4Sb12 with 0.5 wt% Ag additions, the Ag in the composite was determined to react with excess Sb and form an Ag3Sb phase 268 (chapter 6). Examining the binary phase diagram, this Ag3Sb phase likely formed in a eutectic/peritectic reaction during sintering, and a liquid was likely present at 873 K, and possibly at 773 K. This liquid acted as a sintering aid, and the reaction to form the Ag3Sb phase may also act as a beneficial scavenger of excess Sb. The use of Ag as a sintering aid may likely be expanded to other Sb-containing TE materials. Furthermore, the use of a binary or ternary phase diagram between constituent elements of a TE material and a possible addition may indicate materials with a possibly beneficial eutectic/peritectic reactions. 9.3 Elastic moduli of TE materials with nanoparticle or nanoprecipitate additions In general, if the addition of nanoparticles or nanoprecipitates does not change the composition of the matrix TE material, the elastic moduli of the composite exhibits only a weak dependence on the addition, and may be effectively modeled by the Hashin or Halpin-Tsai models. For nanoparticle or nanoprecipitate additions that are typically considered in TE materials (less than 4 vol%), SiCNP additions in Mg2Si (chapter 5) or SnTe (chapter 7), AgNP additions in skutterudites (chapter 6), or precipitates of CdS or ZnS in PbS or PbSe (chapter 8), the elastic moduli are weak functions of the volume fraction of the additions. The weak dependence of the elastic moduli is poorly described by a rule of mixtures approximation (chapters 7 and 8), EC = Vm Em + Vr Er (3) but may be reasonably described by either the Hashin particulate composite [Hashin 1962] (chapters 7 and 8) 269  E V  Er Vr  1  EC  Em  m m  ErVm  Em Vr  1 (4) or the Halpin-Tsai model [Halpin 1992] (chapters 7 and 8)  1  2( a / b )qVr   EC  Em   ErVm  Em Vr  1 (5a) where the boundary condition parameter, q, is  Er    1 Em   q  Er   a     2   Em   b  (5b) These models for the Young’s modulus of the composite, EC, depend on the Young’s modulus of the matrix material and reinforcing material (nanoparticle or nanoprecipitates), Em and Er, and the volume fraction of the matrix and reinforcing materials, Vm and Vr. For the Halpin-Tsai model, the length/thickness ratio, a/b, for roughly equiaxed particles or precipitates was set to 1. 9.4 Fracture toughness and slow crack growth in Mg2Si and other TE materials The fracture toughness, KC, of TE materials is low, typically less than 1.5 MPam1/2, and must be considered carefully when designing at TE module. Despite the importance of KC, it has not been measured for many TE materials. The KC for both unreinforced Mg2Si (chapter 4) and up to 4 vol% SiCNP-reinforced Mg2Si (chapter 5) was measured, as well as for YbAl3 (chapter 3) and Ba0.3Co4Sb12 (chapter 6). Fracture toughness is a function of both the reinforcing material and the powder processing/sintering conditions. Planetary ball milled Mg2Si with addition of SiCNP reached a maximum KC at 1 – 2 vol% SiCNP, increasing 33% over no addition, but 270 vibratory milled Mg2Si exhibited no improvement in KC with 0 – 2 vol% SiCNP addition (chapter 5). The ball milled Mg2Si with 1 – 2 vol% SiCNP may not be the best choice for highest strength , despite having the highest KC, ~ 30% higher than without SiCNP or the vibratory milled material (chapter 5). The fracture stress, σfract, is a function of KC, the flaw size, a, and shape factor, Y, where [Rice 1998] √ (6) When fracture occurs, the critical flaw, acrit, is typically the largest flaw in a region of high stress, and this flaw can be from a variety of sources, such as the pores and cracks that form within a material or scratches and cracks at the surface from machining. Typically, the largest flaw not caused by machining scales with GS. Vibratory milled Mg2Si specimens exhibited an average GS (0.4 – 0.8 µm) roughly 1/5 the average GS of planetary milled specimens (2 – 4 µm). If the critical flaw scales directly with GS, the fracture stress would be ~70% higher with the vibratory milled Mg2Si than the ball milled Mg2Si with 1 – 2 vol% SiCNP , even with the lower KC. If, however, the critical flaw does not scale directly with GS, as might happen with flaws from machining or other external sources, the highest KC would be the more desirable material. In addition, slow crack growth is a particular concern for low fracture toughness materials such as most TE materials. With slow crack growth, a flaw, such as from cutting or grinding to fabricate a leg, may grow to a large flaw that leads to failure in the minutes or hours after machining. Slow crack growth was examined for unreinforced Mg2Si in dry room air and determined to not be a concern. However Mg2Si does react with water and a humid or other environment may remain a concern (chapter 4). 271 REFERENCES 272 REFERENCES [German 1985] German, R.M. Liquid phase sintering. Springer Science+Business Media, Inc., Troy, New York, 1985. [Halpin 1992] Halpin, J.C. Primer on composite materials analysis. Technomic Publishing Company, Inc., Lancaster, Pennsylvania, 1992. [Hashin 1962] Hashin, Z. The Elastic Moduli of Heterogeneous Materials. Journal of Applied Mechanics 29 (1962) 143–150. [Kim, Bush 1999] Kim, H.S., Bush, M.B. The effects of grain size and porosity on the elastic modulus of nanocrystalline materials. Nanostructured Materials 11 (1999) 361–367. [Ni, Ren, Case, Timm 2009a] Ni, J., Ren, F., Case, E., Timm, E. Porosity dependence of elastic moduli in LAST (lead–antimony–silver–tellurium) thermoelectric materials. Materials Chemistry and Physics(2009) . [Ni, Ren, Case, Timm 2009b] Ni, J.E., Ren, F., Case, E.E.D., Timm, E.J.E. Porosity dependence of elastic moduli in LAST (lead–antimony–silver–tellurium) thermoelectric materials. Materials Chemistry and Physics 118 (2009) 459–466. [Rice 1998] Rice, R.W. Porosity of Ceramics. Marcel Dekker, New York, 1998. 273 10 Future work In the examination of the mechanical properties of YbAl3, there were considerable differences between the experimental measurements of elastic moduli and published models, and even between the models themselves (chapter 3). These were stated to possibly be related to “complexities in modeling and understanding the nature of the intermediate valence nature of the Yb ion in YbAl3.” As such, a closer examination of the modeled and measured elastic moduli of other Yb-based materials may be investigated. For Mg2Si, slow crack growth in a dry environment does not occur (chapter 4). However, machining fluids may be water-based or contain dissolved water or other polar molecules required for slow crack growth, and storage of machined components may not be in a dry environment. Additional measurements of Mg2Si and other TE materials in a humid environment or with machining oil on the surface should be investigated to determine if environmental assisted crack growth may occur in other common environments in which the TE materials may be stored. The fracture strength of Mg2Si specimens has shown improvement when SiCNP was added to ball milled powder, but not vibratory milled powder (chapter 5). The difference in the behavior between these two processing techniques may be related to grain size, but requires further exploration. For non-cubic materials, KC may reach a maximum at a critical grain size [Bennison 1989], although this is not shown for cubic materials. The composite thermoelectric (TE) materials may warrant further 274 investigation to determine if an improved model for fracture toughness of cubic matrix materials with addition of nanoparticles may be observed. Additional work on discovery and use of sintering aids can benefit both the cost and development of TE materials (chapter 6). Use of AgNP in other Sb-containing TE materials may enhance sintering, and the use of binary or ternary phase diagrams should be employed to discover other candidate materials for liquid phase sintering aids. The functional relationship between elastic moduli of SnTe1+X as a function of x has only been partially addressed (chapter 7), especially since the stoichiometry of the SnTe is not in the previous elasticity literature [Beattie 1969; Baughman 1969]. The differences between the measured PbS (chapter 8) and the literature values are a concern and should be further explored. The doping effect of contaminants and/or excess Pb on the elastic moduli of PbS is significant in the measured results. 275 REFERENCES 276 REFERENCES [Baughman, Lefever 1969] Baughman, R., Lefever, R. Czochralski encapsulation growth of GeTe, SnTe and PbTe single crystals. Materials Research Bulletin 4 (1969) 721– 726. [Beattie 1969] Beattie, A.G. Temperature Dependence of the Elastic Constants of Tin Telluride. Journal of Applied Physics 40 (1969) 4818. [Bennison, Lawn 1989] Bennison, S.J., Lawn, B.R. Role of interfacial grain-bridging sliding friction in the crack-resistance and strength properties of nontransforming ceramics. Acta Metallurgica 37 (1989) 2659–2671. 277 APPENDICES 278 APPENDIX A. Effect of nanoparticle addition on the elastic modulus of a composite material The effect of the nanoparticle additions on the Young’s modulus, EC, of a nanocomposites material can be predicted from numerous models. In this appendix, we list relationships for four composite models, namely rule of mixtures (ROM), Reuss constant strain (RCS), Hashin particulate (HP) and Halpin-Tsai (HT) [Schmidt 2013]. For each model, the Young’s modulus of the composite, EC, is the based on Er, the elastic modulus of the nanoparticle reinforcement phase, Vr, the volume fraction reinforcement phase, Em, the Young’s modulus of the matrix phase, and Vm, the volume fraction matrix phase. Expressions for the four models can be written as: ROM: EC = Vm Em + Vr Er RCS: (9) V V 1  m  r EC Em Er (10)  E V  Er Vr  1  HP: EC  Em  m m  ErVm  Em Vr  1 (11)  1  2( a / b )qVr   HT: EC  Em   ErVm  Em Vr  1 (12) In the HT model, a/b is the aspect ratio (length/thickness) for the reinforcing phase and q is a boundary condition parameter given by  Er     1 E q  m  Er   a     2  E m   b (13) 279 Table A.1. For spherical metallic particle additions to brittle matrices [Fujieda 2012; Hasselman 1965], comparison of the measured composite modulus, EC, with the values predicted by the four models given in this appendix (rule of mixtures, ROM, Reuss constant strain, RCS, Hashin particulate, HP, and Halpin-Tsai, HT [Schmidt 2013]). Em is the Young’s modulus of the matrix material [Fujieda 2012; Hasselman 1965]and Er is the Young’s modulus of the reinforcing material [Lowrie 1967; Macfarlane 1965; Neighbours 1958; Chang 1966]. Material [Reference] Sodium borosilicate glass + W [Hasselman 1965] Dental porcelain + PtNP [Fujieda 2012] Em (GPa) 80.5 [Hassel man 1965] 80.5 [Hassel man 1965] 80.5 [Hassel man 1965] 80.5 [Hassel man 1965] 80.5 [Hassel man 1965] 60.1 [Fujieda 2012] Dental porcelain + AgNP [Fujieda 2012] 60.1 [Fujieda 2012] Sodium borosilicate glass + W [Hasselman 1965] Sodium borosilicate glass + W [Hasselman 1965] Sodium borosilicate glass + W [Hasselman 1965] Sodium borosilicate glass + W [Hasselman 1965] Er Vr (GPa) (%) 409.8 10 [Lowrie 1967] EC ROM RCS HP (GPa) (GPa) (GPa) (GPa) 90.9 113.4 87.5 92.1 ± 1.4 HT (GPa) 96.2 409.8 20 [Lowrie 1967] 105.5 ± 1.1 146.4 95.9 105.5 114.1 409.8 30 [Lowrie 1967] 118.0 ± 0.2 179.3 106.1 121.1 134.7 409.8 40 [Lowrie 1967] 137.5 ± 0.3 212.2 118.6 139.6 158.7 409.8 50 [Lowrie 1967] 159.9 ± 0.5 245.2 134.6 161.9 186.8 177.6 11.7 [Macfar lane 1965] 82 21.4 [Neigh bours 1958; Chang 1966] 67.2 ± 1.6 73.9 65.2 67.5 69.7 67.8 ± 4.3 64.8 63.7 64.2 65.0 280 Table A.1 summarizes the Young’s modulus change due to the addition of both micron-sized and nanosized-metallic particles to brittle matrices in studies by Hasselman [Hasselman 1965] and by Fujieda et al. [Fujieda 2012]. Table A.1 gives the measured composite modulus, EC, for the addition of 0.10, 0.20, 0.30 and 0.40 volume fraction of tungsten particles with diameters of approximately 30 µm to a borosilicate glass [Hasselman 1965], along with the predicted values of modulus calculated by the ROM, RCS, HP, HT models and the modulus of tungsten [Lowrie 1967]. In addition, Table A.1 lists the measured EC, for composite specimens from Fujieda et al. [Fujieda 2012], who measured the elastic modulus change induced by adding either 26 wt% platinum nanoparticles (PtNP) [Macfarlane 1965] or 26 wt% AgNP [Neighbours 1958; Chang 1966] to a dental porcelain, where the mean diameters of the PtNP and AgNP nanoparticles were 5 nm and 10 nm, respectively. The measured composite moduli, EC, (Table A.1) [Fujieda 2012; Hasselman 1965] agree quite well with the moduli calculated in this study using the Halpin-Tsai model (HT). Since the metal particles were spherical in both the Hasselman [Hasselman 1965] and Fujieda et al. [Fujieda 2012] studies, for the purposes of the calculations we set a/b = 1 in equation 12, where a/b is the particle aspect ratio. The amounts of particle addition in the studies by Fujieda et al. [Fujieda 2012] and Hasselman [Hasselman 1965] were significantly greater than the 0.0068 volume fraction (0.5 wt%) AgNP added in this study. The four composite models given in this Appendix predict a decrease in EC of about 0.39% with the 0.5 wt% AgNP addition to the Ba-skutterudite in this study. In recent research by Schmidt et al. [Schmidt 2013] applied the four modulus-composite models to a thermoelectric system consisting of a brittle matrix and a brittle reinforcing phase, namely with 0.00, 0.01, 0.02, 0.03, and 0.04 281 volume fraction of added SiC nanoparticles (SiCNP) in the brittle thermoelectric matrix SnTe1+X, (where x = 0.0 or 0.016), Hashin and the Halpin-Tsai models best described the elastic modulus data. Thus, for the Hasselman [Hasselman 1965] and Fujieda et al. [Fujieda 2012] studies in which the volume percentage of micro and nano particles added was relatively high (from 0.10 to 0.40 volume fraction), as well as the SnTe1+X (with 0.00 to 0.04 volume fraction SiCNP added), the Haplin-Tsai model agrees with the experimental modulus data relatively well. In particular, for 0.5 wt% AgNP in this study, the Haplin-Tsai model predicts a change in E of only 0.35%, which is consistent with the E values measured in this study for the Ba-skutterudite with and without added AgNP. 282 APPENDIX B. Standard operating procedure for HT-RUS furnace Note: This SOP describes operation of the furnace, not the operation of the RUS computer or software. The furnace controller and power supply was designed and constructed by Brian Wright and Gregg Mulder of the Electrical and Computer Engineering Shop. Precautions Do not operate the furnace with the bell jar removed and the heating element leads exposed. When the bell jar has been removed, if the high voltage is turned on, the line volage leads and the ends of the SiC heating elements are exposed and are are an electrocution harzard. Figure B.1. The top of the furnace has exposed leads to power the heating elements with up to 120 VAC electricity. Do not operate the furnace with these leads exposed. The furnace cools through the outside of the bell jar. When in operation, the bell jar may get extremely hot and cause burns. Additional cooling for the bell jar may be required. The specimen will be exposed to elevated temperatures. Verify that there are no concerns with specimen decomposition or reaction at the temperatures it will be exposed to. 1 Furnace Preparation 1.1 Wear gloves and do not use oils. Oils, especially in the inner furnace, will be burned off during operation. For this reason, do not introduce oils to the furnace. Keep the inner 283 furnace clean, remove any finger oils from the specimen and wear gloves when adjusting furnace components to keep them clean. Figure B.2. Note feedthroughs for the thermocouples (left arrow) and heating element power (right arrow). These feedthroughs and the wires from them should not touch the base or the sides of the bell jar when closed. Figure B.3. Braided insulation must be in place around the conductors to the heating elements to prevent a short. 284 1.2 Clean dust and debris. The furnace is composed of several refractory bricks that let off a significant amount of abrasive dust and debris. For this reason, a thorough cleaning is required prior to operation to ensure the proper sealing and operation of the furnace. Particular attention is required with the feed-through connections for the power and thermocouples, to ensure no shorts are created. Power is supplied through a feed-through with two copper conductors, capable of handling the 20A current to the furnace heating elements (Figure B.3). These copper conductors are shielded with high temperature braided ceramic insulator material, and the insulator must be in place to protect from a potentially dangerous short between the bell jar and the power supply. There are two types of thermocouples, one K-type thermocouple for the lower temperature sensing near the transducers to sense if the transducers are over the maximum design temperature, and a pair of R-type thermocouples to measure the temperature within the furnace. All the thermocouples make connections to the bare wires of the feedthroughs. Figure B.4. The o-ring and the base of the bell jar may have some debris, particularly from the furnace refractories. The debris is abrasive, and may inhibit a proper seal between the bottom and side of the bell jar. The debris should be removed and appropriate grease reapplied to the oring prior to operation. 285 1.3 Clean and reapply vacuum grease. Clean off the centering ring and o-ring at the base of the bell jar (Figure B.4). The o-ring will have Apiezon or equivalent low volatile organic compound (VOC), high temperature vacuum grease on it that may have debris embedded that should be removed. Reapply a thin layer of Apiezon or equivalent low VOC, high temperature vacuum grease. 1.4 Check for electrical shorts. Verify the power connections are properly shielded with braided insulation (Figure B.3) and that they will not be touching the bell jar or any uninsulated components when closed. Turn on the 120 V power for the temperature controllers (NOT the high voltage for the heating elements) and check for room temperature readings from each of the three thermocouples (Figure B.5). Inspect all cables for any damage. Low voltage power switch High voltage power switch Figure B.5. Controls and power supply for furnace. Note the high voltage power supply is activated with the red button, and indicated with a red light. Power for the transducers is provided by a 110 VAC circuit, turned on with a toggle switch and indicated with a green light. 286 1.5 Inspect the hoses. Check the coolant lines to and from the RUS mount assembly (Figure B.6). Check the chiller for proper operation, and to verify the tank is properly filled with water. Turn on the chiller pump to verify no leaks. Turn off after inspection to avoid a buildup of condensation inside the furnace. Inspect the compressed gas hoses for any damage (Figure B.6). Verify the compressed gas cylinder has sufficient pressure to run for the duration of the experiment. Verify the pressure regulator setting does not exceed the pressure rating of the hoses and connectors, typically set much less than 100 psi (for example, 30 psi). 1.6 Set up the bubbler for proper operation. Verify the end of the gas out line that is in the bubbler is below the liquid level (Figure B.7). Adjust the end of the gas out line if necessary to be near the bottom of the bubbler. Refill with water if necessary. Figure B.6. Blue hoses for gas in and gas out (front bottom) and for coolant water in and out (behind valves) should be inspected for damage such as abrasion or holes. Figure B.7. The bubbler (arrow), with the hose extending out, should be filled to cover the end of the hose with liquid to prevent air from entering the hose. 287 1.7 Check the roughing pump. Verify oil is at proper level and is clean. Check for obstructions at inlet and outlet. Note the inlet at the base of the furnace has a screen to catch large debris before it may reach the pump (Figure B.8). 1.8 Check any additional equipment that may be attached or required for the experiment. Figure B.8. Debris can be seen on the screen in the feed-through leading to the vacuum pump. 2 Furnace and sample preparation The sample should be an appropriate specimen for RUS, either parallelepiped or cylinder and with no major chips or flaws. Small chips or flaws may be acceptable. Note the large buffer rods on the RUS transducer will limit the signal available and RUS signal will diminish with increasing temperature, thus a heavier specimen is typically required than for the room temperature setup to ensure sufficient signal. Specimens have been examined with mass of 0.80 g, however a mass of at least 2 g would improve the signal. A higher specimen mass will help keep the RUS signal transferring through the buffer rods. 288 2.1 Measure the specimen mass, dimensions, and determine the thermal expansion for the desired temperature range. The thermal expansion is required to calculate dimension changes for RUS analysis. 2.2 Note any apparent defects on the specimens. 2.3 Verify the RUS transducer rods do not touch the furnace and may be adjusted to the proper angle without hitting the furnace or any other components. Figure B.9. The three transducers are mounted to copper blocks on the chiller plates underneath the furnace. The mount swivels to allow the angle of the transducer to be adjusted. 2.4 DO NOT MOVE TRANSDUCERS BY THE BUFFER ROD. Adjust the RUS transducers to support the edges of the specimen by adjusting the angle of the copper mount that holds the base of the transducer. The mount should allow rotation of the transducer with reasonable hand pressure. If the mount is (a) too difficult to move or (b) does not have sufficient tension to prevent movement, the screw on the swivel mount may be adjusted. Note there is a wave washer to maintain tension on the swivel mount without tightening the screw to fully tight. 289 2.5 Check for RUS signal. Place the specimen on the transducer and run a RUS scan to check for proper signal. Adjust specimen placement and transducer angle to maximize RUS signal. See RUS guide for setup of the RUS software and specimen placement. Figure B.11. Molybdenum baffle being inserted around buffer rods. Figure B.10. Molybdenum sheets to be used as baffles. 2.6 Place molybdenum baffle sheets around the transducers (Figures B.10 and B.11). The specimen may need to be removed during baffle sheet placement. The molybdenum sheets are placed on the second (of three) shelf on the furnace and RUS support, under the insulators that support the heating elements. These should be placed as near as reasonably possible to the RUS transducer rods without touching the transducer rods in order to deflect heat from the furnace before it reaches the transducer base. Verify the RUS signal is maintained. 2.7 Close the front of the furnace with refractories. Fit the front SALI board into position (Figures B.12 and B.13). Fit the front refractory bricks into position (Figures B.14 and B.15). The top of the furnace may need to be slightly lifted to allow the refractories to fit. Verify the RUS signal is maintained. 290 Figure B.12. Front SALI board of inner furnace. Figure B.13. Front SALI board of inner furnace, inserted into position Figure B.14. Front refractory bricks of outer furnace. Figure B.15. Front refractory bricks of outer furnace, inserted into position 291 2.8 DO NOT PLACE HANDS OR BODY UNDER SUSPENDED BELL JAR. Place the bell jar on the furnace. This is a two-person job. Use the hoist to lift the bell jar higher than the furnace, then slide the bell jar into position with the gantry. One person must guide the bell jar by the sides while the second person operates the hoist. Operate the hoist up or down with the remote control wired to the hoist. Guide the bell jar down on the furnace o-ring seal. Bring the hoist down sufficiently that there is some slack in the chain. Verify the RUS signal is maintained. Figure B.16. Hoist and controls to operate up and down. Note the hoist is mounted to a rail for movement left and right, and that the chain has some slack when the bell jar is in position. 292 3 Evacuation of furnace 3.1 Clamp the bell jar to the base with the 12 clamps (Figure B.17) and tighten with a wrench. Figure B.17. Clamps for the bell jar. The same clamps are used to secure the top and base of the bell jar. 3.2 Verify all gas inlet and outlet valves are closed. There are three high vacuum gas valves mounted to the underside of the table, two of which are for inlet and outlet. Close the gas inlet and outlet to snug against the valve seat. The middle valve is for use with a pressure gauge. This middle valve should be left open. Currently, there is no oxygen monitor on the furnace. If a different configuration of the furnace is used that employs an oxygen monitor, then the valve should be closed. 293 Figure B.18. High vacuum valves below the furnace. The gas in and gas out valves should be closed before pumping down the bell jar chamber. 3.3 Pump down the bell jar. Turn on the pump and run until the vacuum chamber has been evacuated, typically about 10 minutes. Verify with pressure gauge. Figure B.19. Flow control panel, with flow controls and bypass for the inert gas in and out of the bell jar and the coolant water from the chiller to the base of the transducers. 3.4 Backfill the bell jar with inert gas. Open the cylinder valve for the compressed gas cylinder containing inert gas and verify the pressure reading on the regulator is less than 100 psi. Verify the gas in bypass valve, located on the flow control panel to the right of the furnace, is closed. Open the high vacuum inlet valve below the furnace. If desired, the chamber may be refilled by opening the refill bypass valve, being careful to avoid disturbing the RUS specimen by the gas. Backfill with inert gas until the chamber is at atmospheric pressure. Verify the RUS signal is maintained. 3.5 Repeat steps 3.3 and 3.4 until at least three inert gas purges of the furnace chamber have been performed. 294 4 Operation of furnace 4.1 Turn on chiller. On the chiller unit, set to appropriate temperature (typically 20°C) and flow (typically 0.5 gal/min). 4.2 Set gas flow. Fill bell jar with gas to slightly above atmospheric pressure to prevent back flow. Open the high vacuum valve below the furnace for outlet gas to the bubbler. Open the gas out valve on the flow control panel. Verify bubble flow through the bubbler. Adjust inlet gas flow control to desired gas flow. 4.3 Set monitor. If an oxygen monitor is attached, open valve to monitor and set oxygen monitor for operation according to manufacturer instructions. (No oxygen monitor is currently attached). 4.4 Set temperature limit controllers. Both temperature controllers are Omega CN3101 series controllers with temperature limits that may be set with the arrow keys. If the controller is tripped, the controller may be reset with the RESET button. The furnace temperature limit controller is typically set to 25° to 50° above the maximum temperature for the experiment. The transducer temperature limit controller is typically set to 80°C. Do not allow the transducers to exceed 80°C to avoid irreparable damage to the transducers or cabling. Figure B.20. From left, the furnace over temperature limit controller, furnace temperature controller, and the transducer over temperature limit controller. 295 Figure B.21. Furnace temperature controller, from left, in run mode, selecting set point 1, and setting set point 1 to 60°C. Figure B.22. 208 VAC power switch with red indicator light on, and 110 VAC power switch with green indicator on. 4.5 Set first desired temperature. Set the furnace temperature controller for the desired first temperature. Use the scroll key on the controller to change to set point 1, use the enter key to select, then use the arrows to choose the desired temperature and enter the selected temperature with the enter key. Use the scroll key to return to run mode. Turn the variac all the way to the left. Turn on the 208/240VAC supply for the heating elements. Increase the variac up to a maximum of 60 V. Verify the amperage on the ammeter to the heating elements is less than 20 A. If the heating rate is insufficient, the variac may be increased up to a maximum of 20 A current. The voltage may be increased up to 120 V for older heating elements until a maximum of 20 A of current. If a higher voltage is required, the heating elements must be replaced (see section 6). Note two set points may be set on the controller at a time, although set point 1 (SP1) is set up to run currently. 296 Figure B.23. Ammeter and voltmeter for heaters. Resistance decreases with temperature, and the current should be monitored to not exceed 20 A. Operate heater elements at less than 60 V for new elements, less than 120 V for elements that have aged due to use. Figure B.24. Heater ammeter, heater voltmeter, and variac output knob on control box. 4.6 Hold at desired temperature. Allow the specimen to soak at the desired temperature until it is uniformly heated, typically 5 minutes soak. Run a RUS scan. Record the settings and time. Do not hold at high temperatures for excessively long time to avoid heating the top of the bell jar excessively. 4.7 Set next desired temperature. Set the temperature controller for the desired next temperature. Verify the amperage to the heaters does not exceed 20 A. Note the resistance of the heating elements changes with temperature so that as the furnace temperature changes, the current through the heating elements changes even when the voltage is fixed. 4.8 Repeat steps 4.6 and 4.7 until RUS scans are gathered for all desired temperatures. 4.9 Turn off 208/240 VAC power to heating elements. Allow furnace to cool below 50°C. Turn off oxygen monitor, if applicable. Turn off gas at regulator. The gas out valve should remain open to allow for excess pressure to escape. Note that the gas will initially have some pressure to continue a flow through the bubbler unit, but as the furnace cools the pressure 297 inside the furnace will drop which could cause the water from the bubbler to be drawn back into the furnace. There is excess hose to reduce the chance of water from the bubbler reaching the furnace, but monitor the bubbler to verify no water reaches the furnace. 5 Opening of furnace 5.1 Turn off chiller and turn off power to the temperature controllers. To prevent a buildup of condensation on the transducers once the furnace is open, the chiller should be turned off just before opening the chamber. To eliminate the chance of electrical energy in the chamber, turn off power to the temperature controllers. Figure B.25. Bell jar at rest on side table and furnace opened. There should be some slack in the chain, allowing all the weight of the bell jar to rest on the table. 5.2 Open chamber. When the bell jar is cooled sufficiently to safely handle, detach the clamps from the bottom of the bell jar with care. Even with the outlet open, there may be a small amount of pressure in the chamber. With one person to guide the bell jar, raise the hoist to lift the bell jar above the top of the furnace. Slide the bell jar to the side table and lower onto the table. 298 5.3 Verify the furnace is cool, then open the furnace. With gloves, remove the front refractory bricks on the furnace and the front SALI inner furnace board. 5.4 Measure the specimen mass and dimensions. 5.5 Note any apparent defects on the specimens. 5.6 Verify all furnace equipment is properly turned off and stowed. 299 6 Replacement of service parts 6.1 Replacement of RUS transducer. If the RUS transducer is no longer providing a proper signal, it may require replacement. Turn all power off. Disconnect the cable from the transducer. Move molybdenum baffle sheets away from the transducer to allow the transducer to be removed. Remove the transducer from the copper mounting block by loosening the clamp screw on the swivel mount, then carefully lift the transducer out of the mounting block. Remove the cold cap from the transducer (Figure B.26). Apply Apeizon or equivalent low VOC vacuum grease to transducer near the top of the brass body to allow heat transfer between the body and the cold cap. Apply a thin layer of grease to the base of the cold cap to allow heat transfer between the cold cap and the swivel mount. Insert the replacement transducer into furnace through the furnace chamber and seat into the swivel mount. Tighten the swivel mount screw. Reconnect the transducer cable. Cold cap on transducer Swivel mount clamp screw Transducer cable Figure B.26. RUS transducer mount components. 300 6.2 Replacement of heating elements. The heating elements should be changed only as a full set, as the resistance of the elements changes as they are used. Do not replace individual heating elements. Disconnect the clamps and the flat braided cables from the four heating elements. Lift the heating elements out of the furnace. Replace with new heating elements. Reconnect the flat braided cable and clamps to the new heating elements. The elements should be wired together in series, supported on the base by an insulating material, and the hot zone should be completely within the inner furnace. Figure B.27. The spiral cut hot zone of the heating elements should be completely contained within the furnace. The power connections to the heating elements should be outside the furnace. 6.3 Replacement of furnace thermocouples. The furnace thermocouples are bare wire type R platinum thermocouples and should only be handled with gloves. Disconnect the old thermocouple from the terminal strip and remove. Thread the new thermocouple through the holes, noting which is the positive and which is the negative side. Connect to the appropriate terminal and trim the excess wire. 301 Figure B.28. Terminal strip on back of furnace for R-type thermocouples, with thermocouple wires extending into the furnace. 302 APPENDIX C. Machine drawings of the HT-RUS equipment Figure C.1. Gantry and full assembly. 303 Figure C.2. Gantry and full assembly BOM 304 Figure C.3. Gantry. 305 Figure C.4. Gantry BOM. 306 Figure C.5. Full assembly. 307 Figure C.6. Full assembly BOM. 308 Figure C.7. Bar clamp. 309 Figure C.8. Base mount weldment. 310 Figure C.9. Bell jar base. 311 Figure C.10. Bell jar base assembly. 312 Figure C.11. Bell jar base assembly BOM. 313 Figure C.12. Bell jar base weldment. 314 Figure C.13. Bell jar base weldment BOM. 315 Figure C.14. Bell jar top. 316 Figure C.15. Cold plate mount 1. 317 Figure C.16. Cold plate mount 2. 318 Figure C.17. Flange tube. 319 Figure C.18. Furnace box. 320 Figure C.19. Furnace box BOM. 321 Figure C.20. Furnace support plate. 322 Figure C.21. Heater support 2. 323 Figure C.22. Heater support. 324 Figure C.23. Inner furnace back SALI board. 325 Figure C.24. Inner furnace base SALI board. 326 Figure C.25. Inner furnace front SALI board. 327 Figure C.26. Inner furnace side SALI board. 328 Figure C.27. Inner furnace top SALI board. 329 Figure C.28. Inner furnace. 330 Figure C.29. Inner furnace BOM. 331 Figure C.30. Outer furnace base center. 332 Figure C.31. Outer furnace base left. 333 Figure C.32. Outer furnace base right. 334 Figure C.33. Outer furnace base. 335 Figure C.34. Outer furnace base BOM. 336 Figure C.35. Outer furnace front 2. 337 Figure C.36. Outer furnace front. 338 Figure C.37. Outer furnace side 2. 339 Figure C.38. Outer furnace side. 340 Figure C.39. Outer furnace top center. 341 Figure C.40. Outer furnace top left. 342 Figure C.41. Outer furnace top right. 343 Figure C.42. Outer furnace top. 344 Figure C.43. Outer furnace top BOM. 345 Figure C.44. Bell jar top. 346 Figure C.45. RUS and furnace. 347 Figure C.46. RUS clamp. 348 Figure C.47. RUS mount assembly. 349 Figure C.48. RUS mount assembly BOM. 350 Figure C.49. RUS mount. 351 Figure C.50. RUS support leg. 352 Figure C.51. RUS support plate. 353 Figure C.52. RUS support. 354 Figure C.53. Splash shield mount. 355 Figure C.54. Splash shield. 356 Figure C.55. Standoff. 357 Figure C.56. Table top. 358 Figure C.57. Transducer cold cap. 359 Figure C.58. Valve assembly. 360 Figure C.59. Valve mount. 361 REFERENCES 362 REFERENCES [Chang, Himmel 1966] Chang, Y.A., Himmel, L. Temperature Dependence of the Elastic Constants of Cu, Ag, and Au above Room Temperature. Journal of Applied Physics 37 (1966) 3567–3572. [Fujieda, Uno, Ishigami, Kurachi, Wakamatsu, Doi 2012] Fujieda, T., Uno, M., Ishigami, H., Kurachi, M., Wakamatsu, N., Doi, Y. Addition of platinum and silver nanoparticles to toughen dental porcelain. Dental Materials Journal 31 (2012) 711– 716. [Hasselman, Fulrath 1965] Hasselman, D., Fulrath, R. Effect of spherical tungsten dispersions on Young’s modulus of a glass. Journal of the American Ceramic Society 48 (1965) 548–549. [Lowrie 1967] Lowrie, R. Single-Crystal Elastic Properties of Tungsten from 24° to 1800°C. Journal of Applied Physics 38 (1967) 4505. [Macfarlane, Rayne, Jones 1965] Macfarlane, R.E., Rayne, J.A., Jones, C.K. Anomalous temperature dependence of shear modulus c44 for platinum. Physics Letters 18 (1965) 91–92. [Neighbours, Alers 1958] Neighbours, J., Alers, G. Elastic constants of silver and gold. Physical Review 885 (1958) 707–712. [Schmidt, Case, Ni, Trejo, Lara-Curzio, Korkosz, Kanatzidis 2013] Schmidt, R.D., Case, E.D., Ni, J.E., Trejo, R.M., Lara-Curzio, E., Korkosz, R.J., Kanatzidis, M.G. Hightemperature elastic moduli of thermoelectric SnTe1±x – y SiC nanoparticulate composites. Journal of Materials Science 48 (2013) 8244–8258. 363