11111111 111111111111 1 I 1 1111111 111‘ 111 l111111111111111111 111111111 11111 . 111111111111111'1111111111 1 111111 111.11.1'1111111 11.11 11111 111 1111111111111111111111111 1111 11111111111 1 1111 111 11.. 1111111111 11111.... .. 111 111111 1111. 1111111 11111111111111.111111'111111 11111111111).1111111111111 1111 111 11111111111 1111 111 .111 1'1 II 11111 1111 11 11" 1111111 .11.11.1 ... I11..1|11 I‘ I.1. II1 I I 1111111111111111111111111111111 11111111 1111111 11 11111 1111111 11 "1.1111111” 111111. 1111111111111 11111 11.111111111111111 1' I II1 1111 ‘3, 1r 1. I. 1‘. F‘ , .11 1111111111 1111 1111 21. "II“ 1 111111 ' , . 1'111 1111111111111111111 111111111111 1111 .1211’11'11191“ 11111 1111 1.111111 11 I 1111111111 111.11 . 111.1. . 111111111111111111 111111111111111111111111111 1.1 ~ 11"" I 1 111111 '1 111111.111111111111111 1111 ’ 1"“ I .I 111 I 1 " E‘s. aII H 1111 1111111111111111 1111111111111.1.1.1111?" . " '4. ‘1 11 11111'.111.111"”11‘ 111111111111111.111.1111: 1...? .1 a... 1‘ 111.11 111-1' - 111151411111 z .3 . - 1i 1111' 11‘1“ 1111 "1’ 11111111111.11111111111111111111111111.‘1‘11'sf'~.1 ‘11. ’ :1. .1'1 .111‘1 7.11: 111.111111111‘11 “51-: 111 1‘1 1‘. 111111111 1.1.1111 11 1111.11111111111n1‘ H11 111 1111 11.. 1“111‘. ‘1” .1111. I 11‘ .11. 1.1“" 11 1111 1111111111 11"”1‘11 11'. 21121.32 '11. ~1..1'111-1111111111111.111 ‘11 1- 1 1111111111... 1111‘1IIIIII 11. 1:311” 311;: 1.111 1' .11 1111. 111. .. 111.11.11.11" 111.111.1111. 1‘ =1 111‘ 11.11111 11111" .7 1111111 11111 111.111.1111 1 1.1 1111‘" . 111... . . IIIIII 1111 111 III 1I1 11111111111 111 111111 ‘1111111.11111L1111111l11.111111111111111 :1 1.1511 111111: 1111111111! 1.11511 “511 :1 I "3E 1111'1 [11111:11‘11‘111'1511111111‘: , 1111' f1: . (”11%; 1.1 1111111111 11.1. 1111‘ I 111 .11... 11111.11 .....1.1 11.111.11.1“ 11;. .. :111-v...-§:1;... '11:??111111‘1'?“1~~-",‘§j 111111 1111 111211' 11‘ 111111111 11111111111111.1111‘11111111111 11111111111111II .1:;I1:1:11111111111§1111114 “‘11::1'1‘111111.1111111111111111;111:1113111‘ 11.1.1'11Il I.|‘. 1 111.1 111 11“ 1111111111111111111111 11! 1:‘11u111...‘ 1111.1 1111 11111111111111.11111111‘1111L111 1...;1. 111111.:1211313131111111 L T 11115111111 1111111 11111 111111111 111111111 11111111111 .11 I11 1111 1111‘ 111.111 1111: 1 1' 111111.111 1111111111111111111. 111 11... III 1‘11'11111 111...I111111 1 II1I111111111I1I1IIII.I1I1 H111 111111 1111111111 1‘. 12.1 11.111111 11:} 11.1.1 1‘ :11 1111111111111 13-1“I 31.1111111911111111111111 11111 11"1 1111 11111111111111 11. 11 111.111 11‘111'11111111111111. 1111111111” "1111111111111 1111111111 1111111 11111111111111.1111 I1I ..... 1’11 111111111111III111111111111111 1:119 111111.In1111 1111 1. .1 111111111111 111111111111 1111111111111111 .11. 111 11111111111111 1111111311111111111111111111111‘111111111111111111171 I 1.1111111. I 111 III 11111111111111 111111 11111111111 .1111. §11111111111111111II11 V 11' 1.11111 1111 111111111111111 1111111111 1'1 ‘111 HH‘1 111111111111 I111 11.1111111111111. 11. ..1 L1'1'111111111111111"1'1111111111 1111111111I1I1111111111 11111111111 1.11111 1 11111 1m1m1 II 11.. 1 1111' ""1111 1'11" 1111 I111 11 111‘ 1“ 1.1 111111111111111111 111m LIBRARY;LI Michiga'ltt. l! University :5 35:; This is to certify that the thesis entitled THE THERMODYNAMICS OF CARBON IN NICKEL-BASED MULTICOMPONENT SOLID SOLUTIONS presented by Daniel Joseph Bradley has been accepted towards fulfillment of the requirements for Ph.D. degree in Chemistry 3644/34“ Major professor [hue December 2, 1977 0-7639 THE THERMODYNAMICS OF CARBON IN NICKEL-BASED MULTICOMPONENT SOLID SOLUTIONS By Daniel Joseph Bradley A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1978 ABSTRACT THE THERMODYNAMICS OF CARBON IN NICKEL—BASED MULTICOMPONENT SOLID SOLUTIONS By Daniel Joseph Bradley The activity coefficient of carbon in nickel, nickel- titanium, nickel-titanium-chromium, nickel-titanium— molybdenum and nickel-titanium-molybdenum—chromium alloys has been measured at 900, 1100 and 1215°C. The results indicate that carbon obeys Henry's Law over the range studied (0-2 at. %). The literature for the nickel- carbon and iron-carbon systems are reviewed and corrected. For the activity of carbon in iron as a function of composition, a new relationship based on re—evaluation of the thermodynamics of the 00/002 equilibrium is proposed. Calculations using this relationship reproduce the data to within 2.5%, but the accuracy of the calibrating standards used by many investigators to analyze for carbon is at best 5%. This explains the lack of agree- ment between the many precise sets of data. The values of the activity coefficient of carbon in the various solid solutions are used to calculate a set Daniel Joseph Bradley of parameters for the Kohler-Kaufman equation. The cal- culations indicate that binary interaction energies are not sufficient to describe the thermodynamics of carbon in some of the nickel-based solid solutions. The results of previous workers for carbon in nickel-iron alloys are completely described by inclusion of ternary terms in the Kohler-Kaufman equation. Most of the carbon in solid solution at high tem- peratures in nickel and nickel-titanium alloys pre- cipitates from solution on quenching in water. The precipitate is composed of very small particles (>2.5 nm) of elemental carbon. The results of some preliminary thermomigration experiments are discussed and recommendations for further work are presented. To Kathleen and My Parents 11 ACKNOWLEDGMENTS I wish to thank the Oak Ridge Associated Universities and Michigan State University for their financial support in the form of a Graduate Participantship and a teaching assistantship, respectively. I would also like to thank Oak Ridge National Laboratory for the use of its facilities and for its financial support during the final term of this work. I am very appreciative of the help I received from the staff of Oak Ridge National Laboratory. In particular, I would like to thank: Mr. James Attril, Dr. James Bentley, Mr. David Braski, Mrs. Sharon Buhl, Mr. O. B. Calvin, Mr. Robert Crouse, Dr. J. C. Franklin, Mr. Greggory Gessel, Dr. Gene Godwin, Dr. William Laing, Mr. John Houston, Dr. Rodney McKee, Mr. Guy Peterson, Mr. Gregory Potter, Mr. Jack Ogle, Mrs. Kay Russell, Dr. James Stiegler, Dr. Robin Williams, Mr. Mark Williams, Mr. Clarence Zachery and Mrs. Joanne Zody. I am sincerely grateful to Dr. James Leitnaker for his direction and encouragementthroughout my time at ORNL. He gave willingly of his time to a sometimes less than thankfulfpupil. My genuine thanks to Dr. Frederick Horne for intro- ducing me to the field of thermodynamics and for encourag— ing my education. Without his counsel and help this work 111 would never have been started or finished. I would also like to thank my fellow graduate students, Dr. Robert Cochran and Mr. Richard Rice for their friend— ship and the many favors they performed for me over the last four years. Finally I would like to thank my family and my wife for their comfort and understanding throughout the long course of formal education. iv TABLE OF CONTENTS Chapter Page .LIST OF TABLES . . . . . . . . . . . . . . . . . . .xvii LIST OF FIGURES . . . . . . . . . . . . . . . . . . . ix I. INTRODUCTION. . l A. Purpose . . . . . . . . . 1 B. Experimental . . . 3 C. Results . . A II. SOLUTION THERMODYNAMICS. . . . . . . . . . . . . 6 A. Chemical Potentials and Activity Coefficients. . . . . . . . . . 6 B. Excess Functions. . . . . . . . . . . . . . . 9 C. Lattice Stabilities . . . . . . . . . . . . . 12 D. Models. . . . . . . . . . . . . . . . . . . . 13 III. CARBURIZATION THERMODYNAMICS. . . . . . . . . . 20 A. Choice of Carburizing Medium. . . . . . . . . 20 B. Analysis of the Thermodynamics of the CO/CO2 Equilibrium. . . . . . . . . . . . 22 C. Analysis of the Literature Data on the Iron-Carbon System. . . . . . . . . . . . 26 IV. ANALYSIS FOR CARBON. . . . . . . . . . . . . . . no A. Introduction. . . . . . . . . . . . . . . . . no B. Procedure for Total Carbon Determination by the Combustion-Gas Chromatographic MCthOd. o o o o o c o o o o o o o o o o o ’41 1. Summary . . . . .'. . . . . . . . . . . . Al 2. Equipment and Reagents. . . . . . . . . . Al Chapter Page 3. Calibration . . . . . . . . . . . . . . . A2 A. Determination of Blank. . . . . a . . . . A2 5. Procedure . . . . . . . . . . . . . . . . A3 C. Precision of the Carbon Analyses. . . . . . . A3 D. Accuracy of the Carbon Analyses . . . . . . . M8 V. EXPERIMENTAL PROCEDURES . . . . . . . . . . . . . 52 A. Preparation of the Alloys . . . . . . . . . . 52 B. Carburization . . . . . . . . . . . . . . . . 59 l. Specimen Preparation. . . . . . . . . . . 59 2. Furnace and Auxiliary Equipment . . . . . 50 a. The Furnace . . . . . . . . . . . . . 60 b. Thermometry . . . . . . . . . . . . . 62 c. Gases . . . . . . . . . . . . . . . . 62 . d. Operating Procedure for the Hydrogen Furnace. . . . . . . . . . . 63 C. Annealing . . . . . . . . . . . . . . . . . . 65 D. Electrolytic Extractions. . . v . . . . . . . 67 1. Description . . . . . . . . . . . . . . . 67 2. Discussion of Precision . . . . . . . . . "69 3. Procedure of Anodic Dissolution of Nickel-Based Alloys for the Concen- tration of Precipitated Carbide Phases. . . . . . . . . . .'. . . . . . . 69 E. Electron Microprobe . . . . . . . . . . . . . 75 1. Introduction. . . . . . . . . . . . . . . 75 2.‘ Procedure for Analysis of Carbide Precipitates. . . . . . . . . . . . . . . 77 vi Chapter Page 3. Calibration Curves. . . . . . . . . . . . 78 F. X—ray Diffraction . . . . . . . . . . . . . .. 84 VI. THE NICKEL-CARBON SYSTEM . . . . . . . . . . . . 88 A. Results of the Carburization Experiments . . . . . . . . . . . . . . . . . 88 B. Comparison with Previous Work . . . . . . . . 96 VII. CARBON PRECIPITATION IN NICKEL AND NICKEL-TITANIUM ALLOYS. . . . . . . . . . . . . 110 A. Discovery of the Carbon Phase . . . . . . . . 110 B. Chemical Analysis of Additional Residues. . . . . . . . . . . . . . . . . . . 112 C. Electron Microscope Results . . . . . . . . . 115 D. Discussion. . . . . . . . . . . . . . . . . . 120 l. Hydrolysis of Dissolved Carbon. . . . . . 120 2. Diffusion Mechanism for Precipitation of Carbon . . . . . . . . . . . . . . . . 121 3. Previous Results. . . . . . . . . . . . . 126 E. Summary . . . . . . . . . . . . . . . . . . . 129 VIII. THE NICKEL-TITANIUM-CARBON SYSTEM. . . . . . . 131 A. Results of the Carburization Experiments . . . . . . . . . . . . . . . . . 131 B. The Solution Thermodynamics of Titanium in Nickel-Titanium Carbon Solid Solutions . . . . . . . . . . . . . . . 1A2 IX. NICKEL-TITANIUM-MOLYBDENUM—CHROMIUM CARBON SYSTEMS . . . . . . . . . . . . . . . . . 1M6 A. Results of the Carburization Experiments . . ... . . . . . . . . . . . . . 1M6 vii Chapter Page B. Carbide Precipitates. . . . . . . . . . . . . 157 1. Carbide Composition . . . . . . . . . . . 157 2. Annealing Experiments . . . . . . . . . ._163 3. Carburization Experiments . . . . . . . . 168 C. An Unidentified Phase of High Carbon Content . . . . . . . . . . . . . . . . . . . 177 )C. THE KOHLER—KAUFMAN EQUATION . . . . 182 A. Calculation of the Nickel-Carbon and Iron Carbon Interaction Energies. . . . . . . 182 B. Analysis of the IroneNickel-Carbon System. . . . . . . . . . . . . . . . . . . . 185 C. Calculation of Interaction Energies in the Nickel-Titanium-Carbon-System . . . . . . 19“ D. Calculation of the Molybdenum—Carbon and Chromium-Carbon Action Energies . . . . . . . 196 E. Prediction of Carbon Solubilities . . . . . . 203 )(I. THERMOMIGRATION... . . . . . . . . . . . . . . . 210 A. Introduction. . . . . . . . . . . . . . . . . 210 B. Experiments . . . . . . . . . . . . . . . . . 212 C. ReSults . . . . . . . . . . . . . . . . . . . 214 XII. SUGGESTIONS FOR FURTHER WORK. . . . . . . . . . 216 A. Analytical Chemistry. . . . . . . . . . . . . 216 B. Experiments . . . . . . . . . . . . . . . . . 217 APPENDIX.......................219 BIBLIOGRAPHY.....................2143 viii Table 301 3.2 3.3 “.2 5.1 5.2 5-3 5.A LIST OF TABLES Thermochemical Data for the CO/CO2 System. . . . . . . . . . . . . . . . . . The Solubility of Graphite in Gamma Iron. . . . . . . . . . . . . . . Date of R. P. Smith (19A6) for Activity of Carbon in y-Iron in Equilibrium with CO/CO2 Gas Mixtures. . . Calibration Data for LECO Gas Chromatograph Carbon Analyzer with National Bureau of Standards Standard Reference Material 121B Analysis of NBS Standards . . . . . . . Fabrication Schedule for Alloys 7261-7268 . . . . . . . . . . . . . . Alloy Compositions as Determined by Several Methods of Analysis, wt % Composition of Alloys Used for Calculations. . . . . . . . . . . . . . . . Results of Multiple Extractions of 0.64 cm Rod Specimens of Ni + 2 wt % Ti + 0.1 wt % C . . . . . . . . . . . . ix Page 27 29 32 A“ 50 56 57 58 7O Table 5.5 5.6 6.3 6.“ 6.5 Analyses of Precipitates by a Colorimetry or Atomic Absorption Spectroscopy and by an Electron Microprobe Energy Dispersive x—ray Analysis . Analysis of Precipitates by Pashen- Runge Emission Spectroscopy and Energy Dispersive x-ray Analysis. Experimental Results for Carburiza- tion of Nickel. Carbon in Nickel. Results of Wada 33 El- Solubility of Carbon in Equilibrium with Graphite (a0 = 1). Results of Smith (1960) for the Activity Coefficient of Carbon in Nickel at lOOO°C. Comparison of Activity Coefficients, Excess Enthalpies, and Excess Entropies of Carbon in Nickel 'The results of Wada 22 a1. (1971 for the (1971) for the Activity Coefficient of Page 79 81 89 101 102 106 108 Table Page 7.1 Results of the Extraction of Alloy 3 (Ni + 1.7 wt % Ti + 0.09 wt % c) Annealed at Temperatures from 1260 to 760°C. . . . . . . . . . . . . . . . . . . 111 7.2 Results of Extraction of Ni—270 and Ni-270 + T1 Alloys Carburized at 1215°C and Then Quenched. . . . . . . . . . . 113 8.1 Experimental Results of the Carburiza- tion of Nickel-Titanium Solutions . . . . . . 132 8.2 Activity Coefficient of Carbon in Nickel-Titanium—Carbon Solutions. . . . . . . 133 8.3 The Temperature Dependence of f0 and the Values of afig and ASE in Nickel-Titanium-Carbon Solutions. . . . . . ..‘ 1A1 8.“ Activity Coefficient of Titanium in I Nickel-Titanium Carbon Solid Solutions in Equilibrium with Graphite and TiC. . . . . 14“ 9.1 Activity Coefficient of Carbon as a Function of Temperature and Composition in Ni—Ti-Mo—Cr-C Solid Solutions. . . . . . . IA? 9.2 Comparison of Equilibrium Concentra— tions of Carbon in Ni-Ti-Mo-Cr-C Solutions . . . . . . . . . . . . . . . . . . 153 xi Table 9.3 9.9 9.5 9.6 9.7 10.1 10.2 10.3 10.“ Page The Results of the Analysis of the Carbide Precipitates by the Electron Microprobe and X-ray Diffraction. . . . . . . 158 Results of the Annealing Experiments. . . . . 16A Solubility of CArbon in Several Nickel-Based Alloys as Determined from Carburization Experiments. . . . . . . . 169 X-ray Diffraction Data on the Un- identified Phase Alloy 7266 A-7783-97 . . . . . . . . . . . . . . . . . . 178 X-ray Diffraction Data on the Un- identified Phase in Alloy 7266-A- 7783-37 . . . . . . . . . . . . . . . . . . . 179 Calculated Values of Nickel-Carbon and Iron-Carbon Interaction Energies, wFCC 13 . . . . . . . . . . . . . . . . Some Relative Lattice Stabilities for 183 Elements of Interest. . . . . . . . . . . . . 18A The Reanalyzed Results of Smith (1960) for the Activity Coefficient of Carbon in Nickel-Iron-Carbon Alloys. . . . . . . . . 186 The Reanalyzed Results of Wada §t_31. (1971) for the Activity Coefficient of Carbon in Nickel-Iron-Carbon Alloys . . ... . 187 xii Table 10.5 10.6 10.7 10.8 10.9 10.10 Interaction Energies Wigc for the Kohler-Kaufman Formalism, wigc = AiJ + 313T + 01JT2 + D13T3. . . . Incorrect Values for the Binary Interaction Parameters, Wigc Calculated with Only Binary Terms. Interaction Energies in kJ-mol‘l Calculated from the Kohler-Kaufman Equation Including Ternary Terms. Comparison of the Value of the Activity Coefficient of Carbon Calculated Using the Kohler-Kaufman Equation and the Value Determined Experimentally. . . . . . . . . . Activities of the Alloying Elements Calculated Using the Kohler-Kaufman Equation,(Eq. 2.30) Comparison of Calculated and Experi- mental Value of the Carbon Activity Where Precipitation of Titanium Carbide Should Start. . . . . . . Composition of Alloys Used for Calculations. . . . . . . . . . . Data From Carburization Experiment A-7603-97 . . . . . . . . . . . . ' xiii Page 198 200 202 20A 205 207 222 223 Table A.” A.5 A.6 A.7 A.8 A.9 Data from Carburization A-7603-105. Data From Carburization A-7603-106. Data From Carburization A-7603-118. Data From Carburization A-7603-121. Data From Carburization A—7603-l23. . . . . Data From A-7783-A Data From A-7783-1N Data From A-7783-15 Data From A-7783-16 Data From A-7783-l7 Data From A-7783-18 Data From A-7783-19 Carburization Carburization Carburization Carburization Carburization Carburization Carburization xiv Experiment Experiment Experiment Experiment Experiment Experiment Experiment Experiment Experiment Experiment Experiment Experiment Page 223 22A 225 226 227 227 228 228 229 229 230 230 A.17 A.18 A.l9 A.20 A.22 Data From A-7783-20 Data From A-7783-21 Data From A-7783-32 Data From A-7783-33 Data From A-7783-35 Data From A-7783-36 Data From A-7783-37 Data From A-7783-38 Data From A-7783-NA Data From A-7783-“5 Data From A-7783-“7 Data From A-7783-98 Carburization Carburization Carburization Carburization Carburization Carburization Carburization Carburization Carburization Carburization Carburization Carburization Xv Experiment Experiment Experiment Experiment Experiment Experiment Experiment Experiment Experiment Experiment Experiment Experiment Page 231 231 232 232 233 233 23A 23A 235 235 236 236 Table A.27 A.28 A.29 A.30 A.3l A.32 A.33 Data From Carburization A-7783-“9 . Data From Carburization A-7783-57 . . . . . . Data From Carburization A-7783-116. Data From Carburization A—7783-l20. Data From Carburization A-7783-l23. Data From Carburization A-7783-125. . . . . Data From Carburization A-7783-l36. xvi Experiment Experiment Experiment Experiment Experiment Experiment Experiment Page .237 237 238 239 290 291 2A2 Figure 3.1 3. 3 LIST OF FIGURES Page The results of Herzberg and Rao (1999) and Snow and Rideal (1929) displayed as A2F"(J)-ABe(J+%) versus J, where the terms have been defined in the text. The relatively random appear- ance of the Snow and Rideal results indicates a lack of internal con- sistency. 'The Herzberg and Rao results, however, produce a smooth curve . . . . . . . 25 The results of Smith (1996) versus Equation (3.6), tnyc vs y0 (y0 = xC/(l-xc). The x's are experimental points, the zeros, 0, are calculated points and the equal signs, =, indi- cates that the calculated and experi- mental values differ by less than l.9%. . . . 3A The results of Ban-Ya 33 a1. (1969, 1970) versus Equation (3.6), in YC vs yC [y0 = xC/(l-xC)]. The x's are experimental points, the zero's, 0, are calculated points and the equal signs, =, indi- cate the calculated and experimental values differ by less than 1.9% . . . . . . . 35 xvii Figure 3.9 3.5 Page The results of Scheil gt 31. (1961) versus Equation (3.6). in YC vs. yC [yC = xC/(l-xC)]. Scheil _t al. performed sets of experiments at constant PgO/PCO2 rather at constant temperature. The x's are experimental points, the zero's, 0, are calculated points and the equal signs, =, indicate the calculated and experimental values differ by less than 1.9%. (POO/P002 = r) . . . . . . . . . . . . 36 The results of Dunwald 32 a1, (1931) vs Equation (3.6), in YC versus yc [yC = xC(l-xC)]. The x's are experimental points, the zero's, 0, are calculated points and the equal signs, =, indicate the calculated and experimental values differ by less than 1.9%. . . . . . . . . . . 37 Calibration data for the LEGO carbon analyzer. The data was taken on three separate days . . . . . . . . . . . . . . . . 97 Optical photomicrographs illustrating the "memory effect" in Ni-2.5 Ti-8 Mo-8 Cr-0.2 C (at. %) (a) as swaged (b) after 1 hour at ll77°C, (c) after xviii Figure 5.2 5.3 5.9 Page 1 hour at 1177°C + 160 hours at 760°C . . . . . . . . . . . . . . . . . . . . 59 Electrolytic extraction equipment constant voltage power supply, Pt tipped forceps, Pt cathode and magnetic stir plate . . . . . . . . . . . . . 73 The calibration curve for the electron microprobe data. The emission spectroscopy results were not used for determining the shape of the line . . . . . . . . . . . . . . 83 Calibration curve for the analysis of M003 and Ti02 mixture for Mo and Ti with the electron microprobe. . . . . . . . . . . . . . . . . . 86 Activity coefficient of carbon in nickel as a function of carbon activity. Dashed lines are ex- trapolations of solid (i.e., ex- perimental) lines . . . . . . . . . . . . . . 92 in Y0 versus l/T for carbon in nickel. The results of Smith (1960) and of Wada, 32 11. (1971) are the corrected results (see text). . . . . . . . . . . . . . 95 xix Figure 6.3 6.9 6.5 7.1 The results of Smith (1960) and Wada 23 .231. (1971) at 100000, as reported, for the activity coefficient of carbon YC in nickel 1s atom per- cent carbon. Our results are represented by the solid line and the error bars. Carbon activities in iron, at 1000°C calculated from data of Smith (1960) and of Banya £3 91. (1971) and Equa- tion (3.6). The dashed line cor- responds to Henry's Law. Note that the departure from Henry's Law does not occur until approximately 2 atom percent carbon is in solution . The recalculated results of Smith (1960) and Wada ££.§l- (1971) for the activity coefficient of carbon. The results were calculated from their raw data and Equation (3.6) (a) Optical micrograph of nickel- 0.l39 wt % C Specimen quenched in water after 38 hours at 1215°C (b) Bright field electron micrograph XX. Page 98 100 109 Figure 7-3 Page of nickel-0.139 wt % C specimen quenched in water after 38 hours at 1215°C . . . . . . . . . . . . . . . . . . 117 (a) Selected area diffraction pattern of a Ni + 0.139 wt % C specimen quenched in water after 38 hours at 1215°C. (b) Dark field electron micro- graph from the area marked by the circle in (a). The average pre- cipitate diameter is approximately 2.5 nm. . . . . . . . . . . . . . . . . . . . 119 Log10 96 (the time required to achieve equilibrium) versus T/K (6 is the time required for the temperature to drop one degree, r is the quench rate and (xc) has sat been defined by Equation (7.9). The intersection of the horizontal lines with the loglo (96) versus T curve is the temperature below which, with the quench rate indicated, equilibrium cannot be maintained by diffusion, e.g., at r = 167 K-sec"1 diffusion can keep xxi Figure Page the system at equilibrium down to 535°C and at r = 16.7 K-sec'l down to 950°C. . . . . . . . . . . . . . . . . . . 129 8.1 Activity coefficient of carbon in nickel—titanium alloys at 900°C. N1 + 2.9 at. % Ti; ONi + 3.6 at. % Ti; top line from Figure 6.1. Note that experimental error is exaggerated in that 9C, rather than fin 9C, is plotted. . . . . . . . . . . . 135 8.2 Activity coefficient of carbon in nickel titanium alloys at 1100°C. N1 + 2.9 at. % Ti, ONi + 3.6 at. % Ti, top line Figure 6.1. Note that experimental is exaggerated in that 90, rather than in 70 is plotted . . . . . . . . . . . . . . . . . . . 137 8.3 Activity coefficient of carbon in nickel titanium alloys at 1215°C. Ni + 2.9 at. 1 Ti, Ni + 3.6 at. Z Ti, top line Figure 6.1. Note that experimental error is exaggerated in that 70, rather than in 90, is plotted . . . . . . . . . . . . . . . . . . . 139 xxii Figure 9.1 9.3 9.9 Page Activity coefficient of carbon in nickel-titanium—molybdenum—chromium »alloys at 1100°C. The lines repre- sent average values. More data are required, in light of the indications in Chapters VI and VIII that Henry's Law is obeyed in Ni-C and Ni-Ti-C alloys, before a least squares fit is Justifiable . . . . . . . . . . . . . . . . . 199 Activity coefficient of carbon in nickel-titanium-molybdenum—chromium alloys at 1215°C. The lines represent average values. More data are required, in light of the indications in Chapters VI and VIII that Henry's Law is obeyed in Ni-C and Ni-Ti-C alloys, before a least squares fit is Justifiable. . . . . . 151 Lattice parameter of the carbide pre- cipitate as a function of Mo/Ti in the carbide. The line was determined by a least squares fit of the data in Table 9.3 . . . . . . . . . . . . . . . . . . 162 (a) The concentration of carbon in alloy A (Ni + 2.6 at. % Ti + 8.9 at. xxiii Figure 9.5 Page % Mo) in equilibrium with the cubic carbide phase as a function of tempera- ture o = 10% of the bulk carbon concen- tration. (b) The concentration of carbon in alloy 999 (N1 + 2.0 at. % Ti + 8.3 at. % Mo + 8.9 at. % Cr) in equilibrium with the cubic carbide phase as a function of temperature. 0 = 10% of ~the bulk carbon concentration . . . . . . . 166 Atom % carbon versus activity of carbon in several nickel—based alloys at 1215°C. The intersection of the two lines, with the same label, is the solu- bility limit of carbon relative to the carbide phase. The lower line represents the solid solution where the slope is 100/90 (xC = Ac/?C). The dashed lines are an extrapolation of the solid solu- tion lines and represent the amount of carbon in solution at any given activity. The upper lines have been fit by least squares to the data from the two phase region, points that xxiv Figure 9.6 9.7 Page diverged from the straight line behavior exhibited near the intersection were ignored . . . . . . . . . . . . . . . . . . . 171 Atom % carbon versus activity of carbon in several nickel-based alloys at 1100°C. The intersection of the two lines, with the same label, is the solubility limit of carbon relative to the carbide phase. The lower line represents the solid solution where the slope is 100/90 (xC = Ac/IC)’ The dashed lines are extrapolations of the solid solution lines and represent the amount of carbon in solid solution at activities exceeding the solubility limit. The upper lines were fit by least squares to the data from the two phase region . . . . 173 Atom % carbon versus activity of carbon in several nickel-based alloys at 900°C. The intersection of the two lines, with the same label, is the solubility limit of carbon relative to the carbide phase. The lower line represents the solid solution where the slope is 100/?C (x0 = Ac/?c). The dashed lines are XXV Figure 9.8 10.1 Page extrapolations of the solid solution lines and represent the amount of carbon in solid solution at activities exceed- ing the solubility limit. The upper ‘ lines were fit by least squares to the data from the two phase region. . . . . . . . 175 Specimen number 7266 A-7603-97 equilibrated at 1215°C at AC = 0.268. Note the needle like precipitates which are characteris- tic of the unidentified phase. The other precipitates are the MC phase . . . . . . . . 181 Comparison of calculated, 0, and experi- mental, X, values of in 9: as a function of XNi in the Ni-Fe-C system. The experimental results are those of Smith (1960) and Wada gt a1. (1971) (see Tables 10.3 and 10.9).. (a) Calculated points determined from Equation (10.9) with the values wNiC and erC taken from the binary results (Table 10.1). (b) Calculated points determined as'a "best fit" of Equation (10.9); the experimental values were the independent variable and wNiC and erC the dependent xxvi Figure 10.2 11.1 . Page variables. wNiFe and erNi were taken from Kaufman and Nesor (1975), Table 10.5. . . . . . . . . . . . . . . . . . . . . 190 Comparison of calculated, 0, and experi- mental, X, values of 2n 7: as a function of XNi in the Ni-Fe—C system. An equal sign, =, indicates that the experimental and calculated values differ by less than 2%. The calcu- lated values differ by less than 2%. The calculated points were determined as a best fit of Equation (10.5) to . experimental results of Smith (1960) and Wada 22.21- (1971) (see Tables 10.3 and 10.9). Values of‘ineNi and wNiFe were taken from Kaufman and Nesor (1975), Table 10.5. . . . . . . . . . . 193 Sample B-6-B at 25x annealed two hours in the Gleeble. The right hand side of photo is the hot end approximately 1300°C. Note decarburization in hot zone. I O O O O O O O O O O O O O O O O O O. O 213 xxvii' 1R CHAPTER I INTRODUCTION A. Purpose Solid solutions are of great technological importance, in particular in alloy metallurgy and semi-conductor manu- facture. Solid solutions are also of considerable theo- retical interest. According to Darken (1967), no general theory for the solution thermodynamics of strongly inter— acting components has been developed. The best theories to date are the regular solution theory of Hildebrand (1927) and the quasi-chemical theory of Herzfeld and Heit— ler (1925) and Scatchard (1931). Regular solution theory does not account for experimentally-observed negative heats of mixing, and neither theory accounts for experimentally- observed asymmetries in the relative exdess Gibbs free energy. A primary purpose of the work reported here was to check the validity of extending to multicomponent solutions the equation proposed by Kohler (1960) for the relative excess Gibbs free energy of mixing for binary solutions GE(re1) = x1x2(xlwl2 + 2.24.21) (1) .u a. .-v ‘6‘: .h- where x1 is mole fraction and $13 is an interaction energy dependent on temperature. Sigworth and Elliott (1979,1976) and Chipman and Brushy (l968)provide extensive lists of references on then— modynamic investigations of multicomponent alloys. However, no attempt has previously been made to use an analytical expression for the integral relative excess Gibbs free energy of the alloys. The experiments reported here provide data that can be used to determine whether interactions of elements in metallic solutions can be described in terms of binary interactions alone. The Kohler equation as modified by Kaufman (1975) requires that the 013 depend only on com- ponents i and 3: If this binary model can be verified, then the number of experiments needed to describe most systems can be reduced dramatically. There are 9.9 x 106- possible elemental quaternary mixtures but only 5.2 x 103 binary mixtures. A second purpose for the work reported here was to obtain quantitative results for the thermodynamics of multi- component solutions by a multi-pronged attack which includes gas phase carburization coupled with electrolytic extrac— tion and analysis of the carbide phases. Such results are essential in attempting to understand the complicated pre- cipitation processes that occur in multicomponent solid solutions. t §- In 4 I. s. ,.H V\ t... v. a.‘ -.. x . q a. a . -\u .9 D n s . A.‘ ' ‘t. . i— v... a . v . .L . . L. t. . 2. n .. .I- nu. us. In. .. ~,» .n a” u g a.‘ .l B. Experimental Paths Because diffusion in solids is both minuscule and slow, experiments to,determine the thermodynamic properties of solid solutions have been both difficult and time consum— ing. In the study of interstitial elements such as carbon, oxygen, and nitrogen in metal matrices, the problem of slow diffusion rates is alleviated by performing experi- ments at relatively high temperatures. All of the tech- niques developed to take advantage of the relatively large mobilities of the interstitial elements rely on equilibrating the system of interest with a system of known-prOperties. The earliest investigations of the solution thermo- dynamics of interstitial elements involved long term annealing. A mixture of known composition is annealed at a fixed temperature until equilibrium is achieved. The sample is then quenched. The microstructure of the quenched material is studied with an optical microscope or other surface analytical techniques. This method is still used in many phase diagram studies (Stover and ' Wulff, 1959). Although useful information is obtained from this type of experiment, quantitative values for thermodynamic functions are not available from it. The method of welded samples employed by Darken (1999) and Golovanenko, §t_§1, (1973) involves welding two samples of different composition. The concentration de- pendence of the activity for the element of interest is known for one of the samples. After equilibrium is achiev— ed, the composition of each half is determined. The activity of the element of interest in the experimental half is set equal to that in the reference half. The method is limited due to the difficulty in obtaining good bonding between dissimilar materials. A third method, used here, involves annealing speci- mens in an atmosphere in which the element of interest has a constant activity (Dunn and McLellan, 1968; Ban-Ya, gt a1., 1969 and 1970). The Specimens thus equilibrate with a bathing medium. Knowledge of the thermodynamics of the bath allows calculation of the equilibrium activity of the element of interest. C. Results The relative partial molar excess Gibbs free energy of carbon in nickel solid solutions have been determined via a gas phase carburization technique, and quantitative methods for the determination of carbon and metal element concentrations in dilute solutions and in the carbide phase have been developed. The data are used to test the ability of the multi-component Kohler equation to describe the solution thermodynamics of nickel alloys. We show that the equation is adequate for our systems, but that a ternary interaction term must be added to describe the Ni—Mo—C and the Ni-CrbC systems. A ternary term is also necessary to describe completely the Fe—Ni-C system. The application of the parameters determined in nickel solid solutions to other solVent systems are checked by comparing literature values for iron-based systems with those determined here. The results obtained for nickel solution are not always applicable to iron solutions. Thermomigration of carbon in nickel-based alloys is discussed in Chapter 11. Appendix A includes all of the data obtained from the carburization.experiments. CHAPTER II SOLUTION THERMODYNAMICS A. Chemical Potentials and Activity Coefficients For every component i in any mixture of n components, the general formula for the chemical potential is '= u: + RT in a i’ i = l,...,n, (2.1) 11i where pi is independent of composition and a1 is the activ- ity. The values of a1 and u: depend upon each other through the reference state and composition variable chosen. For the pure component reference state and mole frac— tion x1 as composition variable, u: +~RT in X “i 1 91, i = 1,...n, . (2.2) O I where u: is the chemical potential of pure component i at the temperature and pressure of interest and where the activity coefficient 91 referred to the pure component has the property 11m 91 = 1, i = l,...,n. (2.3) Another useful reference state is the infinite dilution state. For component 1 as solvent, pi + RT Rn x ui 1 i’ i = 2,...,n, (2.9) with “i = lim(u1 - RT 1n xi), x + 1 1 lim *1 = 1, i = 2,...,n x1 + l (2.5) For the solvent itself, A _ O = m For the solutes, the chemical potential constants u: and u: are related to each other by - 0 Am ui - 111 + RT in Y1, u° = u” + RT 2n 7° (2 7) i 1 ~ 1’ ' where Aco_ A 0: Y1 - lim Y1, 71 lim Y1. (2.8) x 1 + 1 x1 + 1 Moreover, the two types of activity coefficients are related to each other by Yi with = ?1/?:, §i = Yi/Yi, (2'9) y° ?” = l (2 10) i i ' ° an ideal mixture would have ”i - “i + RT 2n x1, i=1,...,n, (2.11) which is valid for all compositions if and only if u: = u: for all components. dilute solutions. The 00 U1=ui+ 111 Thus, Ideality is approached closely in ideal dilute solution is defined by RT in x1, i = 2,...,n, = u; + RT 2n x1. (2.12) 71 = 1 for all the components in the ideal dilute solute. In many cases of practical importance, including most studies of interstitial elements in alloys, the concen- trations of some solutes are so low that 71 = 1. Then the first of Eqs. (2-12) holds for those solutes in the composi- tion range studied. This does not imply, however, that Yi would be unity over the entire composition range. In par- ticular, effective ideality at high dilution does not imply y: = 1. Thus, 9: f 1, and Eqs. (2.7) and (2.12) yield pi = a? + RT in xi + RT in 9:. (2.13) Thus, Ii is a constant(name1y, 9:), for compositions such that Y1 = 1. Note that Eq. (2.13) is a form of Henry's Law since all the composition dependence of ”i resides in the in x1 term; stated otherwise, the activity of component i is directly proportional to its mole fraction for highly dilute solutions. The formulas displayed so far in this section are valid for any homogeneous phase. When two or more phases are in equilibrium, or when two or more crystalline modifications are stable, we designate the phase by a superscript. For example, for a phase a, Eq. (2.2) becomes a _ 00 a “i - “i + RT 2n x1 + RT 2n 91. (2.19) B. Excess Functions For any intensive property y in a mixture, the excess E property y is defined (Scatchard, 1999; Haase, 1971) by 10 y E y - y . (2.15) where y1d is the value of the same property in an ideal mix- ture formed from the same pure components. Thus, by Eqs. (2.2), (2.11) and (2.15) E pi = RT in 91, i = l,...,n. (2.16) Any total molar property 3 is related to the partial molar properties Z1 (32/3n1)T,P n by 9 J?1 IIMS 1x121. (2.17) and the corresponding excess property is given by n zE = z — 219 = 2 x EB, (2.18) _ i 1 1-1 with -E g — -id Equation (2.16) is an example of Eq. (2.19) for the Gibbs free energy since G1 = ui. s, = -(3ui/3T)p,n , H1 = u, + T Si’ (2.20) 11 we have -E .. A A —E _ 2 , A Si - -R tn 71 - RT (Sin Yi/aT)p,nJ’ H1--RT (unfit/amp“!J (2.21) Moreover, for the total excess properties, ._E n A S = -R 2 x1 [in Y1 + (Bin Yi/aznT)p,n J, i-l J E n . H = -RT 2 x (Bin Y /3£nT) , i=1 i i p,nJ _.E n o = RT 121 x1 in yi. (2.22) Note, for completeness, that —-d —o —id Hi = Hi, 81 = E0 (2.23) p. I :13 20 :3 >4 p. The reason for this rather thorough presentation of well-known thermodynamic quantities is that although our experiments are in the dilute solution range, where the infinite dilution standard state and the 71 activity co- efficients are useful, the mixture theories we wish to dis- cuss are cast in terms of the excess functions Just listed. For the excess functions the pure component reference states are required by definition, and therefore so are 12 the 91 activity coefficients. Put another way, although our solutions are dilute enough to be very nearly ideal, the activity coefficients are 293 nearly unity because it is the pure component—based activity coefficients, the 91, that we calculate. Indeed, in most of our experiments the Y1 are all unity and the Ii are all composition independent constants, namely, 9:. The 9: do depend on temperature, however, and we have, Em Hi =-RT(3£n 7: /8£nT)p, 3'?” = -R[2n ‘7: + (Bin flown]. (2.29) C. Lattice Stabilities Suppose that at a given temperature and pressure, pure component i can exist in two stable phases (crystalline modifications) a and 8. Of course only one of these can exist at equilibrium away from a transition point, but instances of supersaturation, supercooling, etc., are plentiful. The relative stability “:80 is defined by pi : ui - ui . (2.25) To see the importance of relative lattice stabilities, consider an alloy which undergoes a phase transition from 13 the a modification to the B modification. Since the mole fractions do not change, the change in Gibbs free energy in the transition is AGQIB = 68—5“ 8 o X xi“i ' Z xil‘i = Z x ( B -u° ) = Z x [u80 - u°° + RT2ni (xn+xi+xj) nil nEZ x w“ [ xixJ _. 2xnxixj ‘- x nxixJ ] J=:#:=1 J Jki xn+x1+xJ xn+x1+xJ (xn +x i+xj)2 n-1 n-2 n—1 2x x x + - 1 15“ (x1 113k) (2.30) i=1 J=l k=1 (x1+xJ+xk) J#i#k k>J To illustrate the physical implications of this model on the chemical potential of a species in a multicomponent system, consider a ternary solution: 17 2 ugasafiga: EWC31[_.:§__§__2+__3._L(1-3X)] = (x3-I-xJ)2 x 3+xJ 2 2 x x x +_z_w MGEJ(J -x3)1-w‘{2(21) X3+XJ X3+XJ X1+X2 X2Xl xl+x 2 - V21 (— —) + 7123 (1- -2x3)xlx2. (2.21) The first two terms in Egg describe the binary interactions of component 1 with the other components in the system. They are due to the heats of solution in the binary mix- tures. The next two terms appear to be independent of component three and show that even if only binary interac- tions are considered all binary interactions affect the chemical potential of a species, not Just the terms involv- ing it. The wg23 term involves ternary interactions, for which there are few data. The wa3 term can be regarded as the extra heat of solution in the ternary over that pre- dicted from a linear combination of the binaries. A non- symmetric function in x1 and x2 would be more appropriate in the cases where the three—one and three-two interactions are appreciably different. In the limit as x3 + 0, Eq. (2.31) yields lim Ed 0 a a 2 a 2 x3 I 0 u3 1p13x1 + W23x2 ‘ *12x1x2 ‘ $21x2x1 (2°32) 18 If the $12 and 931 terms are small compared to the 933 and 713 terms, this equation is analogous to one suggested by Wagner (1952). In the limit of infinite dilution, with l = solvent, Eq. (2.31) yields lim Ea a Am x1 + 0 u3 = $13 = RT in Y3, lim a . a x1 + 1 RT E33 = 2031 - 9w13, lim a a x1 + 1 RT P333 = 3¢31 - 5Wl3, lim Ea x1 + 1 p1 = 0. (2.33) Thus, the Kohler equation reduces to Henry's Law for the solute in the limit of a dilute solution and to Raoult's Law for the solvent. The experiments reported here provide a test of the Kohler formalism and provide data on the solution thermo- E carbon alone cannot lead to all the interaction energies. The dynamics of nickel-based alloys. Measurements of u other ones must be obtained from the literature. The solu— tion thermodynamics of the transition metal binary systems of interest have been determined more extensively and more precisely than have the thermodynamics of these same metals with interstitials such as carbon, oxygen, and 19 nitrogen, so the literature is rich in information on binary metals. If the model is correct, once precise values of wij and wijk, for a system have been determined, then the chemi- cal activity of all the species at all temperatures and composition can be calculated. When the activity data are coupled with thermodynamic data about precipitate phases, the relative stabilities of the various phases can be calculated as well as the phase diagram. CHAPTER III CARBURIZATION THERMODYNAMICS A. The Choice of the Carburizing Medium One of two gaseous equilibria is ordinarily used to control the activity of carbon in solids; namely P80 C02(g) + C(S.S.) : 2CO(g), Kl = —— (3.1) P002% _ PCH + 9 2H (g) + C(S.S.) 1 CH (g), K = . (3.2) 2 9 2 P2 A H2 c Samples are placed in a reaction chamber, at a tempera- ture of interest, together with a gas mixture of known, constant composition.) Knowledge of the value of the equilibrium constant for the gas reaction allows the ac- tivity of the carbon in the sample at equilibrium to be calculated. There are three difficulties with the C02-CO reaction: (1) the amount of CO2 in the mixture becomes very small at high temperatures, which complicates analysis of the gas composition; (2) before it reaches the sample, carbon mon- oxide gas tends to decompose in the furnace to carbon dioxide and amorphous carbon, which causes uncertainty in 20 21 the carbon activity of the gas at the sample surface; and (3) the presence of a small amount of oxygen in the carbon monoxide-carbon dioxide mixture complicates the analysis because of the reaction 2co2 z 200+02. (3.3) The COB-CO reaction is thus unsuitable for use in studies of materials containing stable oxide formers. Although problems one and two have been avoided by most investi- gatOrs, the problem of oxide formation cannot be overcome. Reaction (3.3) controls the oxygen partial pressure, and if an oxide is stable at that pressure it will form. The methane-hydrogen reaction requires a cleaner system, primarily because of the devasting effects small amounts of water or oxygen can have on the gas compositions. Ellis 33 31 (1963) quantified this effect and found that even the addi- tion of a phosphorous pentoxide trap does not eliminate the problem. Bungardt 23 a1. (1969) have shown that results comparable with those obtained from CO/CO2 studies are possible if sufficient care is taken. The advantage of the H2'CH9 reaction is that the oxygen potential can in principle be kept as low as desired. Since titanium and molybdenum are facile oxide formers, the CHu-H2 reaction was used exclusively in this work. 22 Instead of direct analysis of the gas mixture, the carbon content of a pure iron standard was used to determine the carbon activity. The composition-activity relationship for carbon in iron, determined by the CO/CO2 method, has been studied extensively in the past 50 years. Dunwald and Wagner (1931) performed the first quantitative experiments on iron-carbon binaries, and the system has been studied by many others including Smith (1996) and Ban-Ya 32 a1. (1969) and (1970). B. Analysis of the Thermodynamics of the CO/CO2 Equilibrium “It appears superficially that literature data on the iron-carbon system agrees to within 2%. Close examination, however, shows that the apparent agreement is partially a result of using different values for the equilibrium constant for Reaction (3.1). Smith (1996) determined and used a value 10% lower than that employed by Ban-Ya gt_91 (1970). A literature search undertaken to determine the correct value of the equilibrium constant showed that the disagreement results solely from the use of different values of the absolute entropy, 8%, of carbon monoxide. Ban-Ya 32 31. (1970) used values determined by Clayton and Giauque (1932) from data taken by Snow and Rideal (1929). Smith (1996), on the other hand, used a value determined from his own experiments. The JANAF Thermochemical 23 Tables (1971) agree with Smith (1996), while the NBS Series III Tables (1998) used values calculated by Clayton and Giauque (1932). National Bureau of Standards Technical Note TN 270-3 agrees with JANAF for 8298.15’ but no litera- ture reference is given. JANAF uses the value for 8% of carbon-monoxide determined by Belzer and Savedoff (1953) from spectral data of Herzberg and Rao (1999). In order to determine the correct value of 8%, we checked the quality of the two sets of spectral data by a graphical method due to Herzberg (1939). According to Herzberg, a plot of {[A2 F"(J)]-[9 Be (J + %)1} versus J highlights any systematic or random errors in the data {[A2 F"(J)] equals [R(J-l)-P(J+1)], and Be is the equilib- rium rotational constant for a rigid rotor. R and P refer -1 bands of a vibration-rotation to the J = +1 and J band where J is the rotational quantum number.} Figure 3-1 compares the results of Herzberg and Rao (1999) to those of Snow and Rideal (1929). One would expect a smooth curve with a slightly decreasing slope at high J as the centri- 'b fugal distortion constant, D becomes more important. e’ Snow and Rideal (1929) quote a resolution of "at most" 0.1 cm-l, while Herzberg and Rao (1999) claim 0.01 cm'l. Snow and Rideal (1929) do not state an absolute uncertainty, while Herzberg and Rao (1999) claim an uncertainty of lees than 0.03 cm'l. More recent data on carbon monoxide by .o>g:o spooEm m oosposa .Lo>o3on .mpadmog 0mm cam mponnpom one .mocopmfimsoo Hmcgopcfi mo xoma m wmpmofivzw mpadmop Hmopam pcm 30cm on» mo mocmgwoadm Eoocmb zao>fipmaoh one .pxmp on» CH owcfimmp coon o>w£ mason on» when: .h mamuo> AM+hvomziflhvemm< mm oommadmfio Ammmav Hmoofim Ucm 30cm ocm Annmav 0mm Ucm whonnpom mo mpadmmh one .H.m crowns 25 n. o. n o\ O o\ \ \O [I O O .\ O O \ o o \.\ o o V. \o o “\O O O \ o o\0 . o . o\ o No 7 23233911 024 262m o \ 3»; Sea: oi... oz< 0535: o anmTfib 030 In. zmo OJ 0.. (alum'a t -(t‘),.-JZV 26 Rank gt a1. (1961) and Plyler §£_§1. (1955) do not differ significantly from the results of Herzberg and Rao (1999). The absolute entropy of Herzberg and Rao is the one to use. C. Analysis of Literature Data on the Iron-Carbon System The foregoing analysis dictates that the data of Smith (1996) and Ban-Ya gt_a1. (1969, 1970), Scheil gt a1. (1961) and Dunwald and Wagner (1931) be reanalyzed. Table 3-1 contains the thermodynamic quantities used to calculate the equilibrium constant for the CO/CO2 re- action. The data for log K were fit by least squares, with the result, log10 x = A/T + B + C(T) (3.3) :p II -9137 K, 0A = 9.9 K (11 II 9.602, oB = 8.3 x 10’3 9 -1 , CC = 3.38 x 10'5x'1 ‘ -2.272 x 10‘ K C) II The carburization data were fit by a non-linear least squares procedure to a model first suggested by Darken and Smith (1996) 27 Table 3.1. Thermochemical Data for the CO/002 System.a AG°/kJ-mol-1 f Temp. b c d (K) CO(g) C02(c) C(graphite) loglOK 1000 -200.29 -395.92 0.00 0.238 1100 -209.09 -396.05 0.00 1.096 1200 -2l7.77 ’396.15 0.00 1.715 1300 -226.96 -396.23 0.00 2.278 1900 -235.09 -396.29 0.00 2.757 1500 -293.68 -395.39 0.00 3.170 aJANAF Thermochemical Tables, 2nd Ed. (1971). b -l -l -1 o o a 129 J-mol , 0 ° = 0.09 J-mol K . AHf,298.15 8298.15 COAHO = 95 J-mol-l, 08° = 0.09 J-mol'1 K'l. f,298.15 298.15 dC02(g) + C(gr) = 2CO(g). = 0.019 - calculated a loglOK assuming 08° and OAH° are not functions of temperature. f 28 2 P . 10le A0 = 108 E PEA/K] = 10s 70 + 108 ye. (3.9) P 002 log 90 =1; yC + b + d/T. yo = it; = atom ratio. xFe Darken (1996) derived this equation from a statistical model for dilute interstitial alloys. In the model it is assumed that the dissolved carbon is in one of two energy states; namely, it has either no or one carbon atom in a nearest neighbor interstitial position. Although very simple, the model does an adequate Job of predicting the behavior of carbon in binary metallic solutions. The data from the four different investigations were fit separately to Equation (3.9). Table 3.2 contains the solubility of graphite in iron at various temperatures and the standard deviation of the data for each investigation. Also in Table 3.2 are the results of Gurry (1992) for the solubility of graphite in iron at 957 and 1109°C and the extrapolated value of Buckely and Hume-Rothery (1963) for the solubility of graphite in iron at the iron graphite eutectic (1153°C). Statistically, the data of Smith (1996) and of Scheil g£_a1. (1961) fit the model best with the data of Ban-Ya 33 31. (1969, 1970) being almost as good for 29 Table 3.2. The Solubility of Graphite in Gamma Iron. Temperature (°C) Std. Dev. 800 957 1000 1109 1153 Carbon Std. Dev. Investigator at % (O) %a Smithb 3.83 6.01 6.63 8.15 8.87 2.5 Ban-Ya gt g1,b’° 3.69 5.79 6.91 7.92 8.63 3.0 Scheil g§_g1P 3.78 5.78 6.37 7.77 8.91 2.7 Dunwald gt_§1P 3.62 5.99 6.68 8.90 9.22 5.9 Gurry 6.15 8.10 Buckley gt _1 8.98 a o is the root mean square residual error. bSolubility calculated using the investigators published data and the model suggested by Darken (1996). . cBan-Ya 33 31's 1300°c and 1900°c data were ignored. 3O temperatures below 1300°C. If all of the data of Ban-Ya 22 a1. (1969, 1970) are used, the standard deviation Jumps to 7%. Chipman (1972) observed that the 1300 and 1900°C data of Ban-Ya gt_a1 (1970) are in error. Dunwald and Wagner's (1931) data fit the model with a standard devia- tion of 5%. When the values for the solubility of graphite are calculated from each set of data it is obvious that while each set is internally self-consistent, the results do not agree with one another. A systematic error must be present in at least three of the data sets and possibily all four. Smith's (1996) results are the only ones that agree with the graphite equilibration data within two standard deviations over the temperature range 800 to 1153°C. As a result of the systematic deviation among the data sets, it was decided to use only one set of data rather than an average of all the data. Smith's data were chosen for the following reasons: 1. The fit to the model was very good._ 2. He obtained the presently accepted value for the CO/CO2 equilibrium constant using his equipment. 3. Care was used in checking the accuracy of the National Bureau of Standards standard reference material (NBS SRM) used in calibrating his carbon analyzer. 31 9. His data agree closely with the graphite solu— bility data of Gurry (1992) and Buckley and Hume- Rothery (1963)- The equation for the activity of carbon in iron derived from Smith's data is: logloAc = lOglchYc = (a/T)yC + b+d/T + logloyc (3.6) CA = 2.5% C _ 3 2 a - 3.981 X 10 K, 0a = 1.09 x 10 K b = -8 108 x 10"1 o = 1 33 x 10‘2 o , b o d = 2.212 x 103 K, o = 1.69 x 101 K Smith's (1996) published data are tabulated in Table 3.3. The precision of Smith's (1996) data is 2.5%. It is heartening to note that the graphite and the most pre- cise gas phase carburization data agree. The results of Smith, Ban-Ya £E.§l-: Scheil gt a1,, and Dunwald and Wagner are compared with Equation (3.6) in Figures 3.2, 3.3, 3.9 and 3.5. The x's are experimental points, the zeros, 0, are calculated from Equation (3.6) and the equal signs, =, indicate that the calculated and experimental points differ by less than 1.9%. Smith's 32 Table 3.3. Data of R. P. Smith (1996) for Activity of Carbon in y—Iron in Equilibrium with CO/CO2 Gas Mixtures. Carbon T a 2 (°C) wt % at % yc PCO/Pco 800 0.393 1.58 0.0160 2.25 0.356 1.63 0.0166 2.96 0.377 1.73 0.0176 2.65 0.905 1.86 0.0189 2.85 0.993 2.03 0.0207 3.11 0.953 2.07 0.0212 3.12 0.522 2.38 0.0299 3.63 0.568 2.59 0.0266 9.21 0.608 2.77 0.0289 9.50 0.697 2.99 0.0303 9.87 0.661 3.00 0.0390 5.11 0.726 3.29 0.039 5.59 0.726 3.29 0.039 5.69 0.765 3.96 0.0358 6.07 0.815 3.68 0.0382 6.55 0.831 3.75 0.0390 6.75 0.838 3.78 0.0393 6.81 0.836 3.77 0.0392 6.89 0.875 3.99 0.0910 7.29 1000 0.0360 0.168 0.00168 1.98 0.0987 0.226 0.00227 2.99 0.0563 0.261 0.00262 3.12 0.0790 0.393 0.00399 9.21 33 Table 3.3. Continued. Carbon T a 2 (°C) wt % at % yc PCO/PCO 1000 0.133 0.615 0.00618 7.29 0.292 1.115 0.0113 13.8 0.955 2.081 0.0213 27.9 0.655 2.979 0.0307 93.9 0.810 3.658 0.0380 56.2 0.963 9.326 0.0952 70.8 1.081 9.836 0.0508 89.1 1.206 5.371- 0.0568 99.9 1.321 5.860 0.0622 113.3 1.962 6.953 0.0690 130.2 1.966 6.970 0.0692 131.7 1.971 6.991 0.0699 132.9 1200 0.0198 0.0688 0.000688 3.75 0.0191 0.0655 0.000655 3.80 0.0217 0.101 0.00101 5.83 0.0252 0.117 0.00117 7.19 0.0273 0.127 0.00127 7.23 0.0950 0.209 0.00209 12.96 0.109 0.505 0.00507 30.3 0.219 0.992 0.0100 61.9 0.916 1.905 0.0199 122.5 0.913 1.892 0.0193 123.1 0.738 3.391 0.0396 293.6 0.992 9.239 0.0992 352.2 a yo = Xc/XFe - atom ratio of carbon to iron. 39 Ixo vcm UopmHSoHMo map pmsu mmeHccfi .u ompMHSono.mpm .o fimopou mnp .mpsaoa Hmpcmefiamaxm mum m.x one nomv 0% m> o>Ca “Am.mv soapmSUm mamhm> Aw3mav Sufism mo mpHSmmy one .Rm.a can» mmoa an pommfic modaw> Housmefinma \ .mcwfim fiasco mnu cum mosaoa onuflv . ox .m.m mpswfim 3m.m OOH x cm mmmo.o 9 - II 1.1 c... mm .H o - . OOOONHHB n K K r C" u o x x . n m x u . oooooaue m o O n I I an or Ca .8 o on 0 d On x x . - K On I I v D oowflB an n u x 35 .am.a cmnp mmma an poamac mosam> Hancoefipoaxm ocm UopMHzo IHwo on» mmefiocfi .u .wcwfim Hmsoo on» cam mpcfioa UopmHSono who “0 .m.oaou map .mpcfioa Hmpcmefimoaxo mam m.x one .onxIHV\ox u omu om m> 0» Ca .Am.mv COHpmsum mammm> Aowma .mmmav am.mm wwucmm mo mpHSmon one o .m.m mhswfim ooa x a- o.oa wmm.o no o ooooiua on m. S H x 000 o O 00 o O x X! o x x O O x “1 00 x x u OOOOMHIE x x...“ 92.6; x x 1. mo 0 XX X o x x o on u xx ooomfiaue m mm o m 0 mu 0» Ca 0 x o x 0 mm .x x m x o . m 1 O O X K n “0 OOOOOHNB x x x m o m oooomae . m .. mm.m 36 .Aa u moom\ommv .um.a can» wmma hp mommac modam> Hmpcoefipmaxm cam ompmHSOHmo map opMoHocfi .u .mcwfim Hmsvo on» ocm mpcfioa ompmfizoamo mam .o .m.opou on» .mchoa Hmpcme .mpSpmpoQEwp pampmcoo pm ponpmp moom\owm pampw .mfioxnav\ox u emu om Ifipoaxo ohm m.x one Izoo pm mpcmefipoqu go mpom cosmommmm AM.MM Hfimnom .m>.o> Cu .Am.mv coapmzum m5m9o> Aammflv_flm mm afionom mo mpHSmmp mnB .:.m madman 0» m1; mim - - 7 - - 1 - - - v. . m 34 o x o x c c x O o K m x u o o x c . WomwfiuhO O x O O x x .M. o o 7 x o x x o x N x o x o x .. SETH." . c x m x o 3.22” a. > 5 ex 3 m: r. o x on x . 4 K OCX h. mm.:mu& 0 «xx 9 x x o x O MOow .H o 02x .1 .mx 2 Hma 37 .nm.H cos» mmoa an pmmmao mosam> Housmafinmaxm cum oopmHSoHMo on» opmofiocfi .u .mcwfim fiasco on» new mpCHoa ompmasoamo who .0 .w.opmu on» .mpcfioa Hmpcmefipmaxm mam m.x one .mfioxuavox u ozu om m3m9m> 0» nu .Am.mv cofipmsum .m> AHmmHV AM.MM camzzso mo muHSmmp one .m.m maswfim o. OOH x > mmw: . mmmo. - . n . ,. . x a ,9“ mm.a o x OOONOHHB m 00 000 o .3. o o > Cu x x xx x oooooane x a on mo 0 O X 0 I! x x x oooamue u D .x m>.m 38 results scatter uniformly about the calculated points and seem to fit the model in both terms of temperature and composition dependence. The results of Ban-Ya g£_al. for in Y: versus yc, in Figure 3.3 are high compared to Equa- tion (3.6) except at 1150°C, where the results are in better agreement. The residuals at 1150°C, however, are biased as a function of carbon concentration. The results of Ban- Ya g§_§1. at 1150°C were obtained at a different time than those at the other temperatures and this could explain the difference. Figure 3.3 clearly shows that their 1300°C and 1900°C results are not consistent with the model having an intercept which is proportional to 1/T or l/T plus a constant. This affirms Chipman's (1972) assertion that the high temperature data of Ban-Ya gt a1. is in error. The resultscfi‘Scheil gt a; (Figure 3.9), like those of Ban—Ya gt al., are high compared to Equation (3.6). When fit directly to the model, Equation (3.6),Scheil's results do not seem to fit. The residuals indicate that the inter- cept would have to be a complicated function of tempera- ture to fit all the results. Dunwald and Wagner's results are also high compared to Equation (3.6). This is especial- ly true at low carbon concentration where their data indi- cate a zero slope for in Y:‘ Given the precision of the other investigators' results, it is likely that Dunwald 39 and Wagner's results are incorrect at low carbon concentra- tions. Mainly, the results of the other investigators beside Smith were systematically higher for 2n Y: than those of Smith. The most probable reason for this is the gas com- position or the carbon analyses, either of which could conceal a systematic error that would effectively increase the value of the activity coefficient of carbon. Figures 3.2-3.5 all tend to confirm our decision to use only one set of data, that of Smith, in our experiments. If greater accuracy is desired for the iron-carbon system, the areas where improvement of technique would be most valuable are: (1) carbon analyses; (2) analyses of the gas mixtures, and (3) experiments at more, different temperatures to obtain a better fit for the temperature dependence of the activity coefficient. CHAPTER IV ANALYSIS FOR CARBON A. Introduction Analysis for carbon is critical to the results of this work. Considerable effort was expended on develop- ing the combustion method for analysis of carbon and in demonstrating its precision and accuracy. The procedure described here is the culmination of a many step process. The attainable precision of the method is shown to be ap- proximately 1% in Section B; not all analyses were of this precision, however. Section C addresses the question of the accuracy of the analyses. Since the analysis method relies on National Bureau of Standards Standard Reference Materials (NBS SRM), the accuracy of the results depends on the accuracy of the certified analysis of the NBS SRM. Analysis of several NBS SRM's Shows that they scatter approximately 15% relative to their certified concentrations. The scatter in the standards limits the accuracy of the carbon analyses reported here to approxi— mately 15%. 9O 91 B. Procedure for Total Carbon Determination by_the Com- bustion—Gas Chromatographic Method 1. Summary The carbon in the material is converted to carbon di— oxide by combustion in an oxygen stream. The carbon dioxide is then trapped on a zeolite column. After the combustion is completed, the trap is heated, and the carbon dioxide is released into a stream of helium and thence to a chromatographic column. The amount of carbon dioxide is measured in'a thermistor type conductivity cell. The signal is automatically integrated and displayed on a digital panel. The instrument must be calibrated with material of known carbon concentration. 2. Equipment and Reagents Reaction crucibles: fired at 1000°C for eight hours and then stored in a desiccator until used. Acetone: electronic grade, less than 0.0005 percent residue. Tin metal accelerator: washed in water and acetone to remove organic impurities and then dried at 70 to 100°C. Cupric oxide: fired at 1000°C for two to three hours in air. 92 Helium, high purity: passed through a purification train of ascarite, glass wool and Dri-rite. Oxygen, ultrahigh purity: passed through a purifica- tion train of ascarite, glass wool, and Dri-rite. ,3. Calibration NBS Standard Reference Material 121B was used as the calibration standard. Aliquots of less than 20 mg were not used. Homogeneity for aliquots of 20 mg has been demonstrated for National Bureau of Standards Standard Reference Materials (ASTM, E350). 9. Determination of Blank Before actually determining the blank, the instrument is cycled several times with the standard until a constant response is obtained. To determine the blank, one scoop (apprbximately 0.75 gram) of tin granules and then two scoops of cupric oxide are placed in a crucible. The crucible is then placed in the combustion chamber and allowed to sit in the oxygen stream for one to two minutes before cycling the instru- ment. The blank determination is repeated several times until a reading of :1 us is achieved for three consecu- tive determinations. A blank greater than 15 ug indicates 93 that there is probably a leak in the system which must be corrected. 5. Procedure With the instrument stabilized and the average blank determined, the analyses are undertaken according to the following procedure: Each unknown determination is preceded and followed by an aliquot of a SRM. The ali- quots of 121B are measured to contain approximately the same numbers of micrograms of carbon as the samples (1100 pg). Aliquots of standard and sample of less than 100 pg or greater than 1000 pg are avoided. The factor ( ug carbon number of counts ) used for calculating the concentration of carbon in the unknown is obtained by averaging the value obtained for the SRM. .If the instrument is not run for an hour, or if different batches of gas, tin, copper oxide or crucibles are used, the procedure for determining stability and the blank is repeated before proceeding to new samples. C. Precision of the Carbon Analyses Table 9.1 and Figure 9.1 contain data on National Bureau of Standards Standard Reference Material 121B collected in three sets over a period of three weeks. As 99 Table 9.1. Calibration Data for LECO Gas Chromatograph Carbon Analyzer with National Bureau of Stan- dards Standard Reference Material 121B.a NBS SRM Instrument 121B Reading Date (gms) (counts) 2/18/77 0.2213 307.1 0.3029 925.3 0.9395 609.9 0.5209 790.2 0.6091 898.3 0.9157 581.3 0.5192 725.9 0.9080 576.8 0.5182 . 721.6 0.9189 578.5 0.6939 899.0 0.9908 615.0 0.5139 721.7 0.9106 579.3 0.9110 592.0 0.9923 621.1 0.5208 739.1 0.9553 691.9 0.9030 566.2 3/3/77 0.9150 579.7 0.2318 319.9 0.9998 627.0 0.9299 592.8 0.6269 875.0 95 Table 9.1. Continued. NBS SRM Instrument 121B Reading Date (gms) (Counts 3/3/77 0.9550 639.1 0.2219 305.9 0.5113 729.0 0.5965 768.0 0.5089 718.0 0.9296 588.0 3/11/77 0.2927 338.7 0.5513 799.7 0.9302 598.8 0.2116 292.3 0.6199 858.0 0.3369 977.1 0.9158 589.0 0.9273 596.7 0.9159 592.2 0.9520 691.8 0.9199 597.1 0.9358 617.3 0.9085 577.0 0.9322 613.6 0.2277 329.2 aNBS SRM 1213 is stated to contain 0.0720 wt% carbon. 96 coxmp mmz mpmc one .mmmo mumpmamm oops» so .pwnmamcm conme coma map pom wumo coapmspfifiwo .H.: opswam 97 3.208 hzutamhmz. 000 00k 000 00f. 00¢ 00m _ 1 _ _ _ 0 \Nd v.0 .3 I. 0 \1 n.0 Nd v.0 v.0 n.0 «.0 [K 0.0 0.0 n.0 T 0.0 n6 v.0 n no: I. I ah 72:3...v10; «0.54.70: ad a 2:29 . . 2.23 o I so no no n 10:26 a ab 3.58. u '10. u 3‘. +70. x 06 I 2.. 2o ks. \ n \n d nno... to Ian .2568 :79... 9.6.5103. QT I 2.33 hhxgxm 0 I1 0.0 5.0 0.0 _ _ _ _ _ _ 030.05h 030 In. 23.0 (“9015) 952} MS SON 98 shown in the figure the standard deviation in the weight of 121B varied from 3.1 to 9.1 mg. To obtain one percent precision one must use aliquots of 121B with a mass of approximately 0.9 grams or larger. Since 1218 has a nominal carbon concentration of 0.0720 weight percent, aliquots of greater than 300 ug of carbon should be used to ensure one percent precision. In practice it is not desirable to exceed 1000 counts on the instruments. Above 1000 counts the amplifiers begin to saturate and become non-linear in their response. If aliquots of greater than 500 micro- grams of carbon were desirable for some situation a lower amplifier setting can be used, so that the number of counts per microgram of carbon is decreased. D. Accuracy of the Carbon Analyses NBS Standard Reference Materials are used almost uni— versally to standardize instruments for material analysis. These materials undergo a rigorous testing for homogeneity and composition at the Bureau of Standards Laboratory and in private and industrial research laboratories. However, the accuracies of the analyses are not stated or implied by the National Bureau of Standards. The certificate of analysis accompanying the standards shows that in many cases the scatter in the certificate value as reported by the various laboratories is :5 percent for carbon. 99 As part of this research effort several NBS Standard Reference materials with certified carbon contents were examined. Some of the results are tabulated in Table 9.2. The instrument used for these analyses is a LECO carbon analyzer with a gas Chromatograph-thermal conductivity detector. The following procedure was used to measure the carbon concentrations in the NBS materials. The instru- ment was cycled several times until the response stabilized and a constant blank was obtained. A NBS SRM was used to calibrate the instrument. An aliquot of the standard reference material preceded and followed each aliquot of sample. The number of micrograms of carbon was approxi- mately the same in both the calibrating standard and the standard being checked. Table 3.1 shows that the scatter in the data for each standard is less than or equal to 11 percent of the value. The discrepancy with the certified value is as much as :7 percent. The relative lack of accuracy in the certified analysis leads to the following problems: NBS SRM's 1. If one standard is used consistently the precision of results can be greater than 1 percent. The calculated data, however, will contain a systematic error due to the accuracy of the certified analysis. 2. If many different standards are used to calibrate 50 .zonme w .pz omno.o .m Hma 2mm mmz op m>aawfioh mum mCOHumppcmocoo conpmo onem s.o mmao.o mmao.o mmHo.o ozao.o ozao.o maoa m.m- om:o.o mm:o.o mmzo.o emzo.o mzo.o momfi H.mn mmmo.o :mmo.o :mmo.o mmmo.o o:mo.o maoa o.m+ moa.o moa.o moa.o moa.o ooa.o QmH ~.m+ :om.o mom.o zom.o nomqo uma.o mma m.=+ mam.o mmm.o mmm.o mmm.o owm.o mom czam> mmz mwwnm>¢ m m H Apcoosma pnwfiozv monasz Eoph conpmo , 2mm cowpmfi>oa u pmnadz mammamc< mfimzamc¢ mmmno>¢ Am .pzv maoonmu mmeHMprmo .mesmeeMpm mmz co mfimsflmc< .m.: mfinme 51 an instrument, the precision of the measurement will be limited by the scatter in the values of the certified analyses relative to one another. 3. Comparison of data from various investigators is difficult since different groups use different calibrating standards. If different standards are used, discrepancies as high as 10 percent could occur. These problems can be mitigated to some extent if the calibrating standard is cited in the literature. To eradicate the problems, in- accuracies in the standards must be removed. Problems related to inaccuracy have been caused by abuse of the standards rather than by a failure on the part of the NBS. The fact that a scatter of 5 percent is reported on the certification should be sufficient to keep users from claiming accuracies of :1 or 2 percent. Initially, NBS SRM 19E was used to calibrate the instrument and, hence, as a basis for analysis of a num- ber of samples. When SRM 19E was exhausted, SRM 121B was used. All the SRM 1213 data were converted to the SRM 19E after analysis. The correction is shown in Table 9.2. Thus, the data in Appendix A based on SRM 121B were con- verted to the SRM 19E base for all subsequent calculations unless otherwise stated. CHAPTER V EXPERIMENTAL PROCEDURES A. Preparation of the Alloys In working with carbon in alloys containing strong carbide forming elements, special care has to be taken during fabrication. Precipitation of carbides during processing can result in inhomogeneous alloys (Braski and Leitnaker, 1977). The problems of inhomogeneity are not restricted, unfortunately, to the as-fabricated material. It has been found that the carbides cannot be easily removed once formed. The slow diffusion rate of carbide forming metals results in the enrichment of ti- tanium and molybdenum in the former carbide areas even after long anneals. This is demonstrated by the fact that the carbides precipitate in "stringer" like patterns upon aging at temperatures below the solubility limit. Figure 5.1 is an optical photomicrograph showing this so-called "memory effect" in a nickel based alloy similar to those used here. Braski and Leitnaker (1977) concluded that a way to achieve a homogeneous microstructure was to hot work the material at temperatures in the solid solution regime and that any intermediate recovery anneals after cold working 52 53 Figure 5.1. Optical photomicrographs illustrating the "memory effect" in Ni-2.5 Ti-8 Mo—8 Cr-0.2 C (at. Z) (a) as swaged (b) after 1 hour at ll77°C, (c) after 1 hour at 1177°C + 160 hours at 760°C. 59 20 4o soiacnows l0 .0 WI5WX+4F1fl4 NCHES OMS 55 should be in the solid solution regime. As a result of their work the eight primary alloys used in this study were prepared using a slightly modified version of the fabrication schedule suggested by Braski and Leitnaker. Table 5.1 lists the procedure followed. Table 5.2 gives the composition of the melts, as weigh— ed prior to melting, and the composition of the analyzed 3 mm diameter rods. An extra 0.5 weight percent of chromium was added to all of the alloys containing chromium to correct for expected losses through evaporation. The carbon concentration was lowered to one third of its initial value, primarily due to losses during the final deoxidizing anneal. 'As Table 5.2 shows, the molybdenum and the chrom- ium contents of the alloys were analyzed in several dif- ferent ways. Quantitative analysis for transition metal elements in the concentration regimes in which this work was performed is a difficult task due to the high concen- tration of the different elements. The Paschen results for the chromium and molybdenum and the atomic absorption results for molybdenum appear to be unreliable because of the non-reproducibility of these techniques for the elements in question. Table 5.3 gives the values for the composi- tion of the alloys that were Judged to be the best. These values are used in all subsequent calculations. 56 .ummnop some nouns cococoso Loam: who: coEHomam n .AwanV nostpHoH cam memum up omaoHo>oo opsoooopmw 1‘ QOOOHH es s: m new mm ea assess wsfiueeexesa ma H.mm m.m “chapmpmanp Eoopv owmzm oHoo NH oasuHH um :HEan Hmmccm mumHnmEnoch HH a.mm H.= Amszummoasmu Eoopv mwmsm oHoo 0H mossHH um CHEImH Hmmccm opwHooEpoch m m.om H.m msspmpoQEmu Eoopv mwmzm cHoo m oossHH pm :HslmH Hmmccm mumHomEpoucH s m.:: :.w Assaumsanmp Eoopv mwmsm uHoo o oonnHH um sHEImH Hmmccm ouchoEpoch m m.~m w.w Aosspmmean Eoosv owmzm oHoo : coOOMH pm as H Hmmccm MCHuHcoonom m ~.mm m.oH 0 when H.wH =.mH m nmmm m.mm s.MH a sham m.zm m.oH m.mmmm 0.5m m.mH m mmmm mommmo comsuon pmmcms CHEumH m.:m H.mm H mmmm oopsaH es swash uom m =.mmAmmEHu w pHmEmnv ammo nopclpHoE op< H uncoEpmmpe boom ARV AEEV mmmoopm COHumoHsnmm .oz . mchmzm poumEmHo amum wcHhsa com mop< CH coHposomm .mmmhnHmms mmOHH< pom moHsomnom :oHpmoHsnmm .H.m mHan 57 .om u a: «.0 vocHaucoo onHa aOHHa anaz .umu oH coHuaH>oo chaocmuo och .moo o» coauusnsoo pooaav on: can: conuozw .ho um.o mhpxm 0:» onzHOCH nost> .COHUNuHHHuuHO> ho 0» OHQQpSDHLuuc aw wchmwo co anH ucwHoz HHm was» on oHsoz cOHumEonuaaa coow < .uosvoua Hchu 0:» :H oopHnoo ucsosa on» on cocoa no: no um.o .pHoE on» :H mucoEoHo bongo on» o» o>HuaHop .po «0 opsmnopn uona> st: on» on 0:00 .owsau :oHpmhucoocoo anu :H :OHpacHEnouov no you venues chansons unoe on» couooHucoo nH uH .ous> on» no “He no huckupoocs so no: no sou anmana oHuuoesHo>o .mo>p:o :oHuaunHHao any onanopa op can: no: mHOH.HdHhouaE oocououou opaccmuo mmz .mqumou oHuuosHpoHOo on» o» o>HumHon osHm> ann on» you unencuoa mHnwnopn can mwms hoHHa you moon was ano>00ou oann EscoonzHo: .90 new o: 2909 you nonHa> on» 00 am no mucHapnooc: cm was :oHunpouna oHaou on» no um“ u a: m.» 00 uHsnop a o>am no sou anmHocu oHuuoEHnoHooo .ncfiooso so a e: H.o«m.o sooH space m om» «can .AesmHv Ha uo howauHoH own ohsooooua Hao nHumHmcw ho COHuoHaomoc a pom .0: you osHm> season on» go we can .no you osHa> coumpm on» «o uo.m one .HB 900 msHm> powwow on» go u .m zHomeonpaam nHmsco came on» nose acoHuwH>oo Upwccwum my Ho>oH mococHhcoo umo one .woanwcw omozu sou own: no: mpHHHomu mHn» sodomfism o<< . use: no soaonaem .oz conunsusoo afico «cocommm uano acocomwm nQEoo acononmm ”Ho: uooLHa aoHH< wconhwo Encovanoz EsHeouno Echpra .u a: .uHmsz:< ho muocuoz Hmpo>om an cocHEnouoa um ucoHananoo >OHH< .m.m oHan 58 Table 5.3. Compositiona of Alloys Used for Calculations. Alloy Element/wt % Melt No. Ti Cr Mo C N1 7068 2.97 0.087 96.99 7261 2.00 0.015 98.0 7262 1.95 . 12.81 0.019 85.3 7263 2.02 7.78 0.019 90.2 7269 1.95 6.68 0.015 91.9 7265 2.06 9.09 0.016 93.8 7266 1.91 7.08 12.76 0.021 78.2 7267 1.95 3.77 12.93 0.016 81.9 7268 2.00 7.33 6.66 0.015 89.0 7071 2.80 8.08 0.135 89.0 7095 3.06 13.9 0.380 82.7 A 2.00 13.0 0.099 89.9 B 1.73 0.086 98.2 C 2.00 7.90 0.109 90.5 999 1.99 7.36 11.9 0.035 79.0 aThese values were picked from those in Table 5.1. 59 B. Carburization l. Specimen Preparation For the carburization experiments the 3 mm nickel alloy rods were cut into sections 9 cm long. Each specimen was marked with a vibrator tool with the last two digits of its respective melt number prior to cutting from the parent rod. The specimens were then chemically cleaned in a solu- tion of hydrochloric and nitric acid. 'The acid cleaning was followed by washings in methanol and, finally, acetone. After they were cleaned and dried, the specimens were weigh- ed on a Mettler semi-micro balance to 0.002 mg. The samples were next spot welded at each end to loops of nickel wire. It was found that the wire could be re- moved cleanly from the specimens if the welding was done with the proper energy-input (25 watt-sec for 3 mm rod and 0.5 mm wire worked well). If, however, too large an energy—input was used during the welding or if the sample surface became oxidized, then the wire could not be easily removed after carburization. As many as ten samples were welded to the loops in this fashion. The connected set of samples was lowered into the hot zone of the furnace on a nickel tether attached to an iron slug controlled by magnets, as described in Section 2. 2. a. 60 Furnace and Auxiliarnyquipment The Furnace - The carburizing and annealing fur- nace was one of the central pieces of equipment used in this study.. In order to accommodate the wide range of uses required of it, the furnace was designed according to the following criteria: (1) (2) (3) (‘9) (5) (6) (7) It must be capable of being operated safely in an atmosphere of H2 or Ar. It must have incorporated in it a vacuum pump to facilitate changes in sample atmosphere and to check the system for leak tightness. It must allow for cooling rates which vary from a brine quench to a furnace cool. The cooling must be done in an inert atmosphere. It must be inert relative to the gases, e.g. CH“ or H2. Specifically, it must not act as a sink for carbon or a source for any other ele- ments. It must have a constant temperature zone of 9-6 inches. It must allow for reproducible mixing of dif- ferent gases. It must have unobstructed flow of gas around the samples while they are in the hot zone. 61 (8) It must have the capability to purify and monitor the purity of the gas stream. The furnace itself is a platinum resistance heated furnace with a 55 mm bore. The temperature controller used through- out most of the experiments was a Speed-Max G duration ad- Justing (DAT) controller. The controller maintained a constant temperature to 12°C. Toward the end of the in- vestigation an Electromax III current adjusting type con- troller (CAT) was substituted for the DAT. Temperature control of better than 21°C is possible with the CAT controller. ' To insure the inertness of the system the furnace liner is made of DeGussitt-19 recrystallized high-purity alumina. Smith (1996) noted that above 1000°C with a mullite liner the reduction of 8102 becomes a major problem. In this work we found that iron can also be transferred from a mullite liner to samples in a reducing atmosphere. Alumina reduction by hydrogen at the temperatures dealt with here (900-1215°C) is not a problem. The liner is sealed to a copper collar at both ends with a viton O-ring. The water cooled copper collar serves as inlet and exit for gas, as the connection to the vacuum system, and for the removal and introduction of samples. The lower collar contains the vacuum port and connects to the quenching tank through an air—activated 62 gate valve. The upper collar contains the vacuum gage and is fitted with an O-ring groove which allows a pyrex extension tube to be sealed to the collar. The pyrex extension functions as the cold zone of the furnace. A magnet is used to lower the samples into the hot zone. If a quench is de- sired the magnet can be removed and the sample dropped through the gate valve and into the quench tank. The gas system is so arranged that the samples are in a controlled atmosphere until they hit the quenching medium. If slower cooling is desired, the samples can be raised with the magnet into the extension tube. b. Thermometry — The temperature in the furnace was measured with a calibrated platinum-10% rhOdium (Type S) thermocouple. A similar thermocouple was used to control the furnace temperature. Before each set of runs, to insure that the furnace was at the proper temperature, a profile of the furnace temperature was taken. After ini— tially adjusting the resistance across 6 taps the furnace temperature was found to be constant within 2°C over the 100 mm center section of the muffle. No discernible drift in the peak occurred with time. c. Gases - The piping system to the furnace is de- signed to allow three different gas cylinders to be used 63 together or separately. Each of the three lines feeds gas through a Fisher—Porter Tri-flat variable area flow- meter and intozicentral mixing chamber. The flowmeters, with flow rates of 0-300 cc/min, can be used to mix gases to ratios as low as 1:20 with little difficulty. After passing through the mixing chamber the gas stream either enters directly into the furnace or goes through a puri- fication train and then into the furnace. The purification train consists of a palladium catalyst followed by a column filled with Linde 3A molecular sieve. The palladium con- verted any free oxygen in the gas into H20(g) and then the molecular sieve removed the water. The gas stream was analyzed for water on the exit side of the furnace with a Panametrics Model 1000 hygrometer. Water concentrations of less than 0.5 ppm by volume were obtained with this purification technique. d. Operating Procedure for Safe Use of the Hydrogen Furnace 1. Starting 0p a. Close bottom gate valve and unplug electrical Asocket. b. Set all regulators at MS lbs and close all flow meter valves. c. Make sure vent valves are closed. J. 69 Evacuate furnace system with fore pump. (If the fore pump is not used to evacuate the furnace system, a minimum of 0.5 cubic feet of argon must flow through the furnace and more than 1.5 cubic feet is not necessary since the furnace volume is only m.15 cubic feet. Back fill with argon. Repeat d and e for 3 cycles. Open exit valve to exhaust system. The pres- sure in the furnace should be atmospheric or very slightly above. Light pilot light and open exhaust hood. Begin flowing hydrogen with argon still flow— ing. Shut off argon. Shutting Down Start flowing argon. Turn off hydrogen. Flush the furnace with at least 0.5 cubic feet of argon, not more than 1.5 cubic feet is need- ed. (At the end of this time a platinum wire near the pilot light should not be glowing.) Open furnace to remove or insert samples. Leave argon flowing while furnace is open and reclose the furnace as soon as possible. 65 e. Shut off pilot light. f. Shut off argon. g. Close exhaust hood. 3. Use of Quench Tank a. Secure quenching tank to the base of furnace with C-clamps or bolts. b. Flush quench tank with a minimum of 0.7 cubic feet of argon or not more than 2.0 cubic feet. c. Turn off argon first up stream from quenching tank and then down stream Just prior to quench- ing samples. (It is important not to build- up pressure in the tank which may blow the quenching media up into the furnace chamber when the gate valve is opened. d. Plug in gate valve. e. Open gate valve - drop samples into quench tank - close gate valve - unplug gate valve. C. Annealing ,In order to obtain information on the solubility of carbon in the carbide-forming alloys at relatively low temperatures (BOO-1000°C), a procedure other than car- burization was employed. The low solubility of carbon, .<0.05 atom percent, and the slow kinetics of the carburiza- tion reaction make carburization experiments extremely 66 difficult at these temperatures. (See Chapter IX for a discussion of the results of the carburization experiments at 900°C in the carbide forming alloys.) To circumvent the problems of carburization, alloys with a fixed con- centration of carbon were arc melted and cast. Three (3) millimeter rod sections of these alloys were then annealed at various temperatures. For annealing, two platinum wound resistance fur- naces with Inconel 600 furnace tubes were used. The sam- ples were first cleaned as described in Section II B and then wrapped tightly in a sheet of tantalum. The furnaces were designed to allow a continuous flow of argon through the hot zone. The tantalum foil acted as a getter for Ithe impurities in the gas. When samples were being placed in the furnaces the‘flow of argon was increased and was kept high for approximately five minutes after closing the furnace. At the end of an experiment the argon flow was again increased, and the samples were quickly pulled from the hot zone of the furnace and plunged into a 10% sodium-chloride brine. A translucent oxide was visible on alloys containing chromium and molybdenum after quench- ing. Oxidation apparently occurred during the quench rather than during the anneal. The calibrated platinum-10% rhodium thermocouple used in the carburization experiments was used to measure the 67 temperature in the annealing furnaces. The current adjust- ing type of proportional temperature controller was used throughout this series of experiments. The temperature in the region of the furnaces containing the samples was held constant to within 12°C. D. Electrolytic Extractions 1. Description - In order to obtain precise information about precipitated phases in metallic matrices, it is necessary to isolate the precipitate. The precipitate phases in the materials of concern have varied from 0.05 wt z to 5 wt %. Since quantitative determination of weight fractions was desired, a highly specific isolation tech- nique was required such that none of the precipitate phases dissolves but all the matrix dissolves. The literature [Donachie (1972) and Andrews (1966)] indicates that anodic dissolution has been shown to be a highly selective tech- nique. Donachie (1972) lists 9 different precipitate phases that have been successfully isolated by the elec- trolytic technique. Specifically, since MC type phases can be quantitatively recovered and since MC was the phase of primary import in this investigation, it was decided to use anodic dissolution for the concentration of precipi- tates. Anodic dissolution involves using the sample material 68 as the anode and some inert material, such as platinum, as the cathode in an electrolytic cell. The electrolyte most often used, and that used for all this work, is a solution of 10% by volume of concentrated HCl in methanol. Donachie (1972) indicates that in alloys containing tung-y sten, tantalum or niobium a complexing agent such as tar- taric acid must be added to control oxidation since con- siderable amounts of oxides of these elements can form and precipitate. In this connection it was discovered during this investigation that nickel oxide forms in small quantities during electrolytic polishing of surfaces. Oxide formation can be a particularly severe problem in sample preparation for the electron microscope or small angle x-ray scatter— ing. The nickel oxide has only been detected by x—ray diffraction in extracted residue which contained very little MC phase. Since the N10 and MC phases have similar structure and lattice parameters 0.920 nm and 0.931 nm, respectively, the carbide phase, if present, would ob- scure the nickel oxide. That the amount of oxide formed is small is verified by analysis of the extracted material. Nickel varied from a few parts per million to 1000 parts per million but never higher. Another problem cited in the literature [Andrews (1966) Leitnaker (1977)] is the precipitation of silicon in the form of a gelatinous silica during extraction. 69 Leitnaker determined that silica was not precipitating in his high alloy steels with silicon concentration of 1 wt 1. Since the alloys in question here contained only traces of silicon it is certain that, even if it occurred, it would not pose a problem. 2. Precision - In order to insure that the best precision available from the technique was obtained a strict procedure was developed and followed closely in all extractions. (The procedure is outlined at the end of this section.) As a test of the procedure, two samples that had been thoroughly homogenized by long term aging were extracted several times. Table 5.9 contains the results of these extractions. The standard deviation of the procedure is 0.013 wt Z. If 1 gram of material is dissolved, 0.013 wt % corresponds to 0.13 mg. Since each extraction involves the weighing of a centrifuge tube twice with a standard deviation of approximately 0.05 mg, the precision obtained with the following technique is the best that can be expected until a more precise balance and better recovery technique become available. 3. Procedure for Anodic Dissolution of Nickel-Based Alloys for the Concentration of Precipitated Carbide Phases - Equipment and Reagents Semi-microgram balance 70 Table 5.9. Results of Multiple Extractions of 0.69 cm Rod Specimensa of Ni + 2 wt % Ti + 0.1 wt % C. Heat Treatment Quench Precipitate Extracted Temp. Time wt % (°C) (hrs) 1100 16 02° 0.122 0.129 0.103 Avg=0.ll8, o°=0.013 1260 9 Oz 0.128 0.115 0.098 0.108 Avg=0.112, o°=0.013 aThe extraction solution wale% (volume) HCl in methanol. The voltage was held constant 1.5 V for the duration of the experiment «6 hours. bCZ-cold zone cooled. co is the root mean square residual. 71 Constant voltage power supply (0-9 V) Platinum tipped forceps Platinum sheet to serve as a cathode 50 x 70 mm pyrex dish 15 ml centrifuge tube Multi-position centrifuge Ultrasonic cleaner Eye dropper Magnetic stir bar Plastic wrap Methanol-analytical reagent grade Hydrochloric acid-analytical reagent grade Procedure 1. A solution of 10% hydrochloric acid, by volume, in methanol solution is prepared. 2. Any surface oxide is removed from the sample with sand paper. 3. The sample is cleaned by anodically dissolving it for 1 hour. The specimen is held in the platinum tipped forceps which are connected to the positive terminal of the power supply. A piece of platinum sheet functions as the cathode. It is molded to fit the inside of the 50 x 70 mm dish (see Figure 5.2). The dish is filled with the acid solution so that the sample is well covered. Finally, a 72 .opMHQ sHpn oHpocwwE ocm moonpmo um mamohom UoQaHu pm .hHoasw pmzoa owpro> unopmcoo pcoEQHsvo coHpomspxm oHpmHospoon .m.m magma 73 79 piece of plastic wrap is placed over the dish and around the forceps to help control evaporation of the solution. The dissolution is carried out at 1.5 V. The mixture is stirred with a magnetic stir bar. After it is clean, the sample is washed in methanol in the ultrasonic cleaner, dried, and weighed to 0.05 mg. After it is weighed, the sample is placed in a clean dish with fresh solution and dissolved for 6-8 hours as in (3). Care is taken not to get the sample too close to the cathode because the high current that results causes plating on the cathode. A 15 ml centrifuge tube is cleaned with soap, rinsed several times with methanol, and placed in a vacuum dessicator. After 1 hour it is removed and allowed to equilibrate with the air for 1 hour before weighing to the nearest 0.02 mg. Since the precision of the results depends strongly of the precise weighing of the centrifuge tubes in steps 5 and 9, the tube is weighed twice, and 8 the zero is checked both before and after the weighing. The remaining sample is placed in the preweighed centrifuge tube partially filled with methanol, 10. 75 and the tube is then placed in an ultrasonic cleaner to remove any precipitate adhering to the rod. The sample is then removed from the tube, dried, and reweighed. The extraction solution in the dish is transferred to the centrifuge tube with an eye dropper and is spun at high speed for at least 2 min. The supernate is decanted. The precipitate in the tube is washed with methanol and centrifuged again. This procedure is repeated until the supernate is clear. The tube containing the clear precipitate is placed in a vacuum dessicator to remove the methanol. After several hours of dessicating, the tube is allowed to equilibrate with the air for at least 1 hour and is then weighed as in Step 5. If any discoloration or film is visible in the tube, Steps 9 and 10 are repeated. E. Electron Microprobe 1. Introduction The electron microprobe was used to analyze the car- bide precipitates extracted from the nickel matrix. The microprobe offers several advantages over conventional techniques such as atomic absorption spectroscopy or gravimetric analysis. The more conventional techniques 76 usually require large samples, are destructive, and re- quire equipment that was not readily available for this work. Besides requiring only small samples and being non- destructive, the microprobe permits a rapid analysis which is important when substantial numbers of samples need to be analyzed. A method requiring only a small amount of sample was important in this work because often only 1 mg of material was available and several different types of analysis were desired. Abdel-Gawad (1966) and E. W. White 33 31. (1966) have shown that the electron microprobe can be used to analyze quantitatively micro—crystalline powders. The procedure used in this study is essentially that described in their papers. The assumption is that the intensity ratios for elements in the powders are constants for any given com- position. A series of powders was analyzed by conven- tional techniques and then by the microprobe. A calibra- tion chart was then constructed comparing intensity ratios of elements of interest to weight percent ratios. The use of intensity ratios and calibration curves severely restricts the applicability of this technique. Light elements are not detected by the instrument. The calibra- tion curves are complicated with only three elements if a wide range of concentrations are considered. Fortu- nately, the system of interest here is essentially a two component mixture of titanium and molybdenum. Chromium 77 and nickel are also present, but amount to only 1.0 and 0.05 wt %, respectively, and were not considered in the calibration curve. Practically, one is limited to the analysis of, at most, three elements of mass greater than sodium. The instrument used in this investigation was a Materials Analysis Corporation electron microprobe coupled with a Si(Li) energy dispersive x-ray detector and a multichannel analyzer. 2. Procedure for Analysis of Carbide Precipitates To obtain quantitative results from the microprobe a substrate of atomic number less than 11 is necessary. Elements above sodium emit x-rays that are detectable with the energy dispersive x-ray detector, and there is also a greater chance of absorption and fluorescence inter- actions between the substrate and the sample at high atomic number. Beryllium appears to be the best material for our purposes. It has a low atomic number (four) and is available in a sheet form that can be mounted in epoxy and polished to a high sheen. Another requirement of the substrate is that it be an electronic conductor because the surface charge that could otherwise result would lead to erroneous results. The precipitates were dispersed in methanol and then transferred onto the beryllium chip with a Pasteur 78 pipette. The crystallites adhered to the surface of the polished beryllium after the methanol evaporated. It was not necessary to further bind them to the surface with glue or graphite. A constant accelerating voltage of 25 keV was used for the electrons. The beam was caused to raster over an area of approximately 10,000 02. A window of 0.3 eV was ordinarily used for each elemental peak. The peaks normally used corresponded to the Ka of titanium and the La of molybdenum. In a typical analysis the specimen was counted for 20 seconds (N10,000 counts) in ten different locations on the substrate.. The resultant intensity ratios were then averaged. It was also part of the pro- cedure to check for inhomogeneity in the sample by analyz- ing very small areas but no gross inhomogeneity was dis- covered. 3. Calibration Curve Several different carbide precipitates were analyzed by atomic absorption spectroscopy and with the microprobe. The calibration curve was based on materials of very similar composition and crystal structure to the precipitates. Tables 5.5 and 5.6 and Figure 5.3 are the result of this effort. The data were fit by least squares to Intensity M0 (L0) .. ls .M_o2 Tntensity Ti_(KE) - 0.006+0.980 (wt % T1)-0.016 (wt % Ti 79 Table 5.5. Analyses of Precipitates by a Colorimetry or Atomic Absorption Spectroscopy and by an Electron MicroprObe Energy Dispersive X-ray Analysis. d Microprobea Titaniumb Molybdenumc %%—%4%% IMO/IT1 (wt 7) (wt %) 7263e 37.65 A-7783-17 + 8 Cr 7269f 0.91 92.99 91.37 0.97 A-7783-l7 + 9 Mo 7262e 1.27 38.80 98.55 1.25 A-7783-l9 + 8 Mo 7266e 1.98 36.06 52.91 1.97 A-7783-19 + 8 Mo + 8 Cr 7267e 1.33 37.17 99.96 1.39 A-7783-l9 + 8 Mo + 9 Cr 72689:f 0.85 99.35 37.60 0.85 A-7783-19 + 9 Mo + 8 Cr 7266e 3.19 63.1 17.9 3.62 A-7783-37 + 8 Mo + 8 Cr aThe intensity ratio is the average of approximately ten measurements. The root mean square residual is m¢2%. The precipitates were dispersed on a Be wafer to facilitate the analysis. bThese analyses were performed by a colorimetric method. The uncertainty is %5% of the value. 80 Table 5.5. Continued. 0These analyses were performed by atomic absorption spec- troscopy. The uncertainty is m:5%. The weight percent ratio is based on atomic absorption results. th % Mo= 7% wt 5 Ti ° 8The base composition is Ni + 2.5 at. % T1. The additions of the Mo and Cr are in atomic percent of the uncarburized alloy. dBy error analysis fThe chemical analysis of this precipitate was performed at a later date than the others in this table. 81 Table 5.6. Analysis of Precipitates by Pashen-Runge Emis— sion Spectroscopy and Energy Dispersive X-ray Analysis. Microprobea Titaniumb Molybdenumb wt % Moc IMO/1T1 (wt %) (wt %) Wt % Ti 7262 A-7603-97 1.50 30 50 1.66 + 8 Mo 7266 A-7603-97 2.88 72 25 3.00 + 8 Mo + 8 Cr 7095 A-7603-106 1.79 32 ' 67 2.09 + 1.2 Ti + 8 Mo . aThe intensity ratio is the average of approximately ten measurements. The root mean square residual is approxi- mately 2%. The precipitates were dispersed on a Be wafer to facilitate the analysis. bThe root mean square residual is approximately 10%. 0 wt % Mo _ 19%. CBy error analysis _WE_%_T1 - 82 .osHH one so oamnm on» wchHEpoooo pom pom: pom who: prSnos zqoomonpooum COHm anam one .mumc oposQOHoHE coppoon map pom o>pSo COHpmhnHHmo one . m . m 6.33%. . 8935 8.63424 10 ol». 0:5. .5; m N . _ _ _ A :0.» Ev 90.0 I 22er N C. Es .3 . . I . as A2). .e Ev com o + moo o I 03 o. I \o. \— 91 I 543552.548 7: >aoomomeomam zoiamommq 0.202 -22 £5.55 33.5.25 .25 E 82.533 0.5;. e. .3 AT I 1.7\ >aoowomkomam _ ZO.mm:2m Nel> >m ouzémmfio 0.2: $3 7+7 _ _ _ mmnmlhs QBOIJZKO LLI ow] OLLVU ALISNBLNI 89 The root mean square residual is 2%. Initially it was hoped that a calibration curve could be prepared by intimately mixing pure materials such as titanium and molybdenum powders or titanium and molybdenum oxides. Figure 5.9 shows the result of mixing molybdenum oxide (M003) and titanium oxide (T102). A straight line relationship was obtained between intensity ratio (I(Mo)/ I(Ti)) and weight percent ratio (wt T Mo/wt % Ti) however when this result was applied to carbides of a known composi- tion the calibration curve disagreed with the atomic ab- sorption results by a factor of two. F. X-ray Diffraction Precipitates were examined by x-ray diffraction as follows: The precipitates were first dispersed in methanol. The suspension was then dropped onto a glass slide and the methanol allowed to evaporate. The dried precipitate was scrapped off the slide and placed on a silicon single, crystal wafer. The wafer acts as a substrate in the dif- fractometer and is oriented so that silicon diffraction peaks were not detected. A small amount of TaC powder, a0 = 0.995587 1 0.000020 nm, was then sprinkled on the wafer as an internal standard. Finally, a drop of poly- vinyl alcohol was used as a binder. A diffracted beam graphite monochromator rejected all wavelengths except 85 .mnosqosoHE cospomHm on» an3 HR 6cm 0: Mom chaprE NOHB cum oooz mo memecm on» now o>pzo QOHpmanHmo \ .:.m ohstm 86 O.N m; 9:252 No: I nee: z. i .8 E 0.. 22 oxo ts 0.0 _ _ H vmnmlhh 030....sz “I owl OllVl-J AiISNBlNI 87' those corresponding to the copper Kc lines. The scan speed was usually 0.25°/min. A typical experiment ran from 20 to 80° 20. CHAPTER VI THE NICKEL-CARBON SYSTEM A. Results of the Carburization Experiments Appendix A contains a precis of all carburization ex- periments. Table 6.1 contains a summary of the results of these experiments for the nickel-carbon system. In each experiment several specimens were carburized along with an iron standard. Carbon activities relative to graphite, were calculated from Eq. (3.6). The data set numbers in Appendix A and in Table 6.1 refer to the Oak Ridge National Laboratory notebook page numbers where the experiments were recorded. The activity coefficients in Table 6.1 were obtained by dividing the activities by the respective atom frac— tions. As Figure 6.1 shows, the activity coefficients scatter uniformly about a constant value at each of the three experimental temperatures. Calculated slopes were of the same magnitude or smaller than the uncertainties. Thus, the activity is proportional to the atom fraction for these experiments - Henry's Law is obeyed. Solute- solute interactions are therefore negligible or of the same magnitude as solvent-solute interactions for the concentra- tions studied. 88 89 Table 6.1. Experimental Results for Carburization of Nickel. (a) Temp. Carbon A Date Set Ac (°C) (at. %) Y A-7603-106 0.291 1215 0.729 39.9 A-7603-106b 0.291 1215 0.719 90.8 A-7783-37 0.295 1215 0.739 39.9 A-7783-38 0.198 1215 0.938 95.2 A-7783—116 0.138 1215 0.332 91.6 9—7783-120 0.288 1215 0.676 92.6 9-7783—123b 0.270 1215 0.618 93.7 A—7783-9 0.113 1100 0.177 63.8 A-7783-l9 0 129 1100 0.186 69.9 A-7783—15 0.295 1100 0.998 65.0 A-7783-17 0.709 1100 1.05 67.5 9-7783-18 0.596 1100 0.869 62.8 9-7783-19 0.212 1100 0.359 59.8 9—7783-35 0.922 1100 0.661 63.8 A-7783-32L 0.520 1100 0.816 63.7 A-7783-32Hb 0.520 1100 0.816 63.7 A-7783-33 0.967 1100 0.739 63.2 9.7783-125b 0.197 1100 0.303 65.0 A-7783-99 0.601 900 0.999 139 A—7783-95 0.356 900 0.257 139 A-7783-95b 0.356 900 0.259 190 A-7783-97 0.278 900 0.201 138 A-7783-98 0.256 900 0.200 128 A-7783-99 0.110 900 0.0782 191 A-7783-57 0.187 900 0.191 133 A-7783-136b 0.291 900 0.211 138 90 Table 6.1. Continued aActivity of carbon relative to graphite, calculated from Eq. (3.6). The concentration of carbon in iron for each data set is given in Appendix A. NBS SRM 19E is the analytical basis for the above data. quuilibrium reached by decarburization. 91 .mocHH HmpcoEHpono ..o.HV oHHom mo mCOHpmHoqmnpxo ohm nocHH omnmmo .sz>Hpom consmo mo COHpocsm m mm meoHc CH connmo mo ucoHongooo sz>Hpo< .H.w opsmHm 92 zommHpow map pom .coppoqms as .0083 so :33 Hm Mm moo: one 833 £26 so 3392 one .m.m osstm A ex. .3. zomm<0 mm; 00.. 050 00.0 mN.0 0 _ — _ mm . O Eb: .o z 34; o .- 88: 1:55 0 m0 v o I. u o J 2. m. IA 3 0 3 J.- H o O 3 o m lfill 00 w w o a o I. m o 1103 mi» .b. \\ W. o m [I L: O l. 00 O «a o _ nNnn . lb» 030 74210 Figure 6.9. 99 Carbon activities in iron, at 1000°C, cal- culated from data of Smith (1960) and of Banya 23 a1. (1971) and Equation (3.6). The dashed line corresponds to Henry's Law. Note that the departure from Henry's Law does not occur until approximately 2 atom percent carbon is in solution. ACTIVITY OF CARBON ac 0.75 0.50 0.2 5 100 ORNL-DWG 7 7—1114 5 I I I T—UA I 1000°C, K=135.9 BAN-YA er a/. (1970) SMITH (1946) O A o A / / o / A / / A / / / A / / / —I / o I 1 I J 2 4 6 8 at. 7° CARBON 101 Table 6.2. The results of Wada 33 a1. (1971) for the Activity Coefficient of Carbon in Nickel. Uncorrected Correcteda Temp. At % C, At % C, A A °C in Fe in Ni AC Yc Ac Yc 800 3.96 0.963 1.03 222 1,092 225 2.60b 0.259 0.583 230 0.596 235 1000 5.93 0.777 0.693 89.2 0.733 99.3 3.09 0.919 0.307 79.2 0.333 80.9 3.09 0.929 0.307 72.9 0.333 78.5 2.91 0.919 0.291 70.3 0.315 76.1 2.01b 0.210 0.185 88.1 0.201 95.7 1.93 0.178 0.129 69.7 0.136 76.9 1200 5.99 0.97 0.999 95.8 0.960 97.9 3.57 0.608 0.215 35.9 0.223 37.6 1.11 0.122 0.0592 99.9 0.0589 98.3 aActivities recalculated using Equation 3.6 which corrects for the CO/CO2 equilibrium constant. quuilibrated starting from higher carbon content. 102 Table 6.3. Results of Wada gp.a1. (1971) for the Solubility of Carbon in Equilibrium with Graphite (ac=l). _l Temp. At. 7 C <°c) in/Ni, c 850 0.589b 171 1000a 1.07C 93.5 1.09° 91.7 1.02b 98.0' 1.02b 98.0 1.11b 90.1 1.11b 90.1 1.11b 90.1 1.07b 93.5 1197 1.87C 53.3 l.83° 59.6 aMeasured at 997°C and corrected to 1000°C. bSpecimens were packed with graphite powder in an alumina boat. cCarburized by a controlled CHu-H2 mixture with a graphite boat. 103 .Aw.mv coHpmsom ocm mpmo 3mg LHonp song oopmHSonolwfiwm msznos 0:9 .consmo no pcoHonmmoo sz>Huom on» now AHNmHV .Hm no momB osmfiommHv :pHEm mo mszmop canthonsoo.H one .m.m opszm 109 so .30 200.53 mm; 00.. 000 00.0 0N0 _ .. _ :50: ._03 4043 0 .000: 1.52m 0 v3.03 mi... ._ _ _ 0. Nmnn Th» 030 l.. 220 00 nb 00 mm 0.000» 1v woeavo so lwanousaoo All/\IlOV ' 95 105 The uncertainty in the slope (1270) is 0 = 670, and the root-mean-square residual in Io calculated from Equation (6.9) is = 6.2. For atom fractions greater than 0.001, activity coefficients calculated from Equation (6.9) are the same as the one calculated from Equation (6.1) within the mutual experimental uncertainties. Table 6.9 contains the results of Smith (1960) and the values of the activity of carbon calculated using Equation (3.6). Some of the values for the activity of carbon listed in Table l of Smith (1960) cannot be calculated from his Equation 1, even after Equation 1 is corrected for the obvious typographical error. Equation 1 of Smith (1960) should read, with N1 2 xi, log Y2= 10g [(a2)(N1/N2)] = 3.37 (Nz/Nl), (6.5) where the activity coefficient of carbon is relative to the infinite dilution state of carbon in iron. The ac- tivity coefficient 9c relative to graphite is calculated from [see Equation (2.10)] ?c = Y/Ysat.5 likewise, the corresponding activity Ac is calculated from Ac 2 a2/ a2,sat.‘ The "uncorrected" entries in Table 6.9 are calculated from Smith's Table I, which itself contains two incorrect entries: (1) for 6.61 carbon atomic percent in Fe, Smith 106 Table 6.9. Results of Smith (1960) for the Activity Co- efficient of Carbon in Nickel at 1000°C. Uncorrected Correcteda At. % C At. % C, A A in Fe in N1 Ac Yc Ac Yc 1.29 0.192 0.0979 68.9 0.116 81.7 2.75 0.331 0.250 75.5 0.293 88.5 9.99 0.632 0.979 75.8 0.557 88.1 6.19 0.970 0.816 89.1 0.897 92.5 aActivity calculated using Equations (3.6) and the raw data of Smith (1960). 107 reports 0.191 for a2 whereas Equation (6.5) gives a2 = 0.123; (2) for 6.19 carbon atomic percent in Fe, Smith reports 0.115 for a2 whereas Equation (6.5) gives a2 0.110. The recalculated, corrected results of Smith (1960) were fit by least squares to Yc = 82.0 + 1100 Xc‘ The standard deviation of the slope 1100 is 210.. The root mean square residual of Ic is o = 1.9. All of Smith's recalculated results, except one point, lie within 10 of our interpolated results as shown in Figure 6.5. Schenck, gt al., (1965) did not report their raw data, and, although precise recalculatiOn of their results was therefore impossible, they reported Henry's Law be- havior up to the saturation limit of carbon. It is clear from Table 6.5, however, that their results differ from those reported here by about 15%. After analysis of all available nickel-carbon data, we conclude that Henry's Law is obeyed within the pre- cision of the data. The present results and the report of Schenck 33 a1, (1965) indicate the validity of Henry's Law. IThe corrected results of Smith (1960) and Wada, pp a1. (1971) show a slight dependence of activity coef- ‘ficient for any particular composition agrees within 108 Table 6.5. Comparison of Activity Coefficients,a Excess Enthalpies, and Excess Entropies of Carbon in Nickel. Investigator Temperature/°C Sgth?g e8u21.b e8a81.d Bradleye Smithf 7c, 900 119 102 138 136 9C, 1000 76.1 70.5 88.3 89.6 87.7 9c, 1100 53.9 51.9 65.0 69.3 90, 1215 38.3 37.7 96.3 92.0 AHE/kJomol-l 50.2 96 50.3 59 ASE/J-mol'l-K’l 3.9 0.97 1.9 5.0 aCalculated using iron standards and Equation (3.6). bNo estimate of the error was stated by the author. The graphite and the CH9/H2 carburization techniques were used. 1 -1 C0Y=9.5%, 0H=l.0 kJ-mol'l, os=o.08 J-mol‘ °K The graphite carburization technique was used. 0Y=9.2%, oH=3.5 kJ'mol'l, os=2.7 J°mol-1K'l. The graphite and the CHu/H2 carburization techniques were used. d eoy=l.9, oH=3.3 kJ-mol’l, os=2.9 J mol'lx‘l, the CHu/HZ carburization technique was used. on=2.2%, the CO/002 carburization technique was used. 109 experimental error with the corrected results of Smith (1960) and of Wada, et a1. (1971). Table 6.5 is a comparison of the average value of Yo obtained by five investigators. To obtain the value of Y0 at non-experimental temperatures the average values of 70 were fit by least squares to Equation (6.1). Table 6.5 E c and ASE calculated from also contains the values of AH these fits to the data. Dunn and McLellan (1968) have the largest set of data E and ASE from which AHc c have been calculated, and it is apparent from the small size of the uncertainty in their values for the excess functions that their data are internally consistent. However, their activity results are quite different from ours and from those of Wada, g; a;. (1971) and Smith (1960). The differences are out- side the experimental uncertainties of the various sets of data. It appears likely, then, that Dunn and McLellan (1968) have a systematic error in their data. CHAPTER VII CARBON PRECIPITATION IN NICKEL AND NICKEL-TITANIUM ALLOYS A. Discovery of the Carbon Phase In the course of some of the aging experiments describ- ed in Chapter V, electrolytic extraction of specimens of alloy B(Ni + 1.7 wt % Ti + 0.09 wt % C) yielded a black residue which we attributed initially to the presence of titanium carbide in the specimens. This inference was contrary to the Stover and Wulff (1959) nickel-titanium- carbon phase diagram, which showed that the specimens could contain neither titanium carbide nor graphite. Thorough examination of the residue revealed: (1) The residue had a lower density than that of titanium carbide; (2) the residue lacked the characteristic metallic appearance of titanium carbide; (3) x-ray experiments on the residue gave diffraction patterns of much lower in- tensity than patterns from similar quantities of titanium carbide, and the lines were shifted to higher 28 values. (9) Table 7.1 shows that the concentration of the residue is not a function of temperature, whereas the solubility of most carbides in metals increases rapidly as a function of temperature. Clearly, the residue was not titanium 110 111 Table 7.1. Results of the Extraction of Alloy B (Ni+l.7 wt % Ti + 0.09 wt % C) Annealed at Tempera- tures from 1260 to 760°C. Bulka ' Sample Annealing Carbon Precipitate Alloy Number Temp./°C Time/hrs. (wt %) (wt %) B B-15 1260 16 0.08 0.19 B—15A 760 168 0.08 0.16 B A-7609 1100 16 0.09 0.12 A-7609 1260 9 0.09 0.11 8Specimens were analyzed after aging. 112 carbide. Some remaining possibilities for the residue are: (a) It is not present in the alloy specimen but is instead a product of the extraction process; (b) it is free carbon that has precipitated from solution during quenching; (c) it is an amorphous phase produced by precipitation of alloy impurities such as oxides and sulphides. B. Chemical Analysis of Additional Residues New alloys containing only small concentrations of carbon were prepared. The carbon content was adjusted to any desired level by annealing the specimens in CHu/H2 mixtures. The low carbon concentrations provided an easy check of possibility (c) above and also provided homo- geneous materials which could be examined by electron microscopy. The results of the electrolytic extraction of the gas- carburized alloys are presented in Table 7.2 along with the analyses of the extracted residue for carbon. Some observations on and inferences from the table are: (1) No measurable residue is collected from uncarburized nickel. That is, no carbon means no residue, and possibility num- ber (c) above is eliminated. (2) The residue is approxi- ‘mately 96 to 75 wt. % carbon. (3) Most of the carbon, both in the nickel and in the nickel—titanium alloys, is 211.3 .r .ba 5H0.0 mH second on» conga coHusHon an conhwo,uo unsoew owmpo>w 029 a“ naught oh w>HuHmuom mH minus: 0 p mach .coHpmNHpsnpao now on» o» aoan o a .ps 00.0? cocHaucoo 000p zOHH Hoo I up . cununmh ohwswm cwofi poop on» mH .um0.0Io .mspagmaaw zuH>Huozccoo Hugues» 00mg a co vocHEQOuoo aw: conuaom 00HIm005I< - «9 a .um 000.0: mmH.o 0.00 mwu.0 muH00.0 mmmmw.0 03H.0 0000.0 comp< 0m w.m Hz HwHImoo~I¢ \ I H9 a .um HH0.0 snw0.0 m.wm hno.0 00000.0 0520.0 mmn0.0 mm0.o coup< ma a.NIHz om.o omHoo.o momo=.o smmewwmu 000.0 JHH.u m.mw H.0 00H00.0 00mmo.H 0:H.0 :NH.0 Lona: 0m a.wIH2 0m.0 m0H00.0 mahmm.0 so mno.0 0mm.“ 0.ms Hm.0 0mH00.0 00H00.0 Imomuna 0m.0 me00.0 H0000.0 mMH.0 mmH.0 poumx 0m x - nm.0 m0000.0 mwmzm.0 omH.0 HH.0 oao.o aun.o e.mm mHm.o «afloo.o 0Hm~0.0 Asa.o H=H.o uoms< em we“ Imuosua Hz mUu.0I :m0.0 a.m~ mn0.0 00000.0 00nam.0 0m0.o H40.o cowu< 0: HmH Imomsu< Hz mmH moooo.l mmwhh.o omoo.o .oom Imowmt< n< H2 cm.Hcoum couuooom osounom ARV HwVI va nHmmHmc< owcmnu mucosa “~20 onH< M0 w .u3 00 v .u3 MC V .u3 cm QUHHOU ONCGEU hn u.u3 #IvaS OEdE wuunwm 0H .ruhmom . pHmom ucwfioz “tonnwu coEHoonu wouooHHoo EH coEHoouw L02 cunhmu nouumo btoHuowpuxm coHpeNHssnsmo HH 0H m m h w m a m N H .omuuuwnc cock 0cm oomHmH um owuHhsanu mmoHH< «B + OHNIHZ 0cm ONNIHZ ho COHuocLuxm 80 m uHfimmm .m.» mHnmh 119 recovered in the residue. The amount of carbon not ob- tained as a residue from the electrolytic extraction of the quenched alloys is 0.01710.019 wt % and there is no statistically significant difference in the specimens with and without titanium. (9) From column 11, the concentra- tion of carbon remaining in solution after the quench is slightly higher in the water quenched specimens. However,' the difference is probably not significant because of ex- perimental uncertainty and the small number of experi- ments. (5) Chlorine analysis and metal analysis on both of the A-7603-97 alloys gave a metal to chlorine atom ratio of 3 to 5. The chlorine contamination is a result of the extraction procedure. The precipitates were dif- ficult to separate from the supernates due to their low densities. There is little doubt that the chlorine is present in the form of nickel and titanium chlorides, and that if the chlorides were absent only carbon would re- main. The non-reproducibility of chlorine is related to the scatter in column 9. (6) X-ray experiments on the residue yielded extremely weak, unidentified diffraction patterns in the case of the residues from the nickel— titanium alloys and no diffraction at all in the residue extracted from samples of carburized nickel. 115 C. Electron Microscgpe Results Examination of the quenched specimens in the electron microscope did not clarify the nature of the residue. In bright field the matrix of the specimens appeared to be one phase (Figure 7.1). Selected area diffraction revealed the presence of a second phase in both alloys (Figure 7.2). However, the phase indexed as face centered cubic with a lattice parameter aO N 0.92 nm, the same as nickel oxide. Coatings of oxide have been recognized in other nickel- based alloys (Kenik and Carpenter, 1977).' Stereoscopic examination of the micrographs did not place the precipi- tates conclusively. While it seemed clear that many were on the surface, some particles appeared to one of the three observers to be within the foil. Attempts to adjust the sample preparation technique to avoid oxide formation proved fruitless. The electron microscope work indicates only that if a precipitate phase is responsible for the residue, then the precipitates are smaller than the 2.5 nm diameter particles shown in Figure 7.2. Small angle x-ray scattering experiments undertaken to determine whether precipitates exist in the alloy matrix also failed to yield conclusive results, for the same reason yi§., scattering of the nickel oxide layer on the surface of the specimens. 116 Figure 7.1 (a) Optical micrograph of nickel-0.139 wt % C specimen quenched in water after 38 hours at 1215°C. (b) Bright field electron micrograph of nickel-0.139 wt % C specimen quenched in water after 38 hours at 1215°C. 117 20 40 60 MICRONS I O 140 500x 0.001 INCHES 0.005 Figure 7.2 (a) (b) 118 Selected area diffraction pattern of a Ni + 0.139 wt % C specimen quenched in water after 38 hours at 1215%. Dark field electron micrograph from the area marked by the circle in (a). The average precipitate diameter is approxi- mately 2.5 nm. 119 120 D. Discussion We have shown that a carbon residue is electrolytically extracted from quenched specimens of nickel and nickel- titanium initially at 900 to 1200°C. The cooling rates used have no measurable effect on the amount of carbon precipitated, and all but 0.017 wt % is in the residue. There remain two possible explanations for the behavior (1) the isolated carbon atoms in the matrix form the resi- due during the electrolytic extraction process, or (2) the carbon is precipitating from solution during the quench. Hydrolysis experiments, discussed in the next paragraph, show that the extracted residue is carbon that precipitates during the quench. l. Hydrolysis of Dissolved Carbon Hydrolysis experiments on heavy metal carbides (not alloys) by Bradley, Pattengill and Ferris (1965) and Ferris and Bradley (1965) have shown that carbides hy- drolyze to form methane and other alkanes in basic and neutral aqueous solutions and to form carbon dioxide and organic acids in acidic solutions. The authors state that they have no experimental evidence to suggest that graphite forms, during the hydrolysis, and moreover think graphite formation unlikely because radicals such as HCO, :CO and CH2 form instead of graphite. 121 The nickel-carbon solid solutions studied here are essentially substoichiometric carbides with even less carbon-carbon bonding than in the carbides discussed by Ferris and Bradley (1965). If the carbon in our samples were in solid solution, the hydrolysis experiments indicate 'that the individual carbon atoms would be oxidized to carbon dioxide. On the other hand, if the carbon is present in the alloy specimens as an elemental phase, then the extrac- tion process would not affect it. Since the extraction _ experiments resulted in the isolation of a carbon residue, the carbon must not have been in solid solution; i.e., the carbon precipitated during the quench. 2. Diffusion Mechanism for Precipitation of Carbon In this section we show that the diffusion rate of carbon is fast enough to account for the observed agglom- eration during the time of cooling.8 Diffusion is a strong function of temperature. Smith (1966) reported that the diffusivity, D, of carbon in nickel varies with absolute temperature, T, according to D = 0.366 e-17’900/T cm2 sec-1 (7.1) During diffusion carbon atoms migrate from solution at t/0 a rate proportional to e- (deGroot, 1951), where the 122 relaxation time, 0, is given by e = d2/n20, ' (7.2) with d the distance over which diffusion occurs. Dif- fusion is 99% complete when t Z 90. In the precipitation experiments under discussion here, the specimens were cooled at a rate of approximately 170 K sec"l (Beck and Bigot, 1965). The specimen temperature thus decreases by one degree in about 6 milliseconds. When 90 is smaller than 6 msec, the diffusion process is fast enough to be completed during the time interval required for a one degree temperature decrease. When 90 is larger than 6 msec, the diffusion process is too slow to be com- pleted during the time interval, and precipitation begins to cease. When the temperature falls low enough that 90 is very large compared to 6 msec, carbon atoms diffuso so slowly that no further precipitation is observable. Figure 7.3 is a plot of 90 versus absolute tempera- ture on the assumption that the diffusion path length is 10 nm. This estimate is based on Figure 7.2 where any carbon particle cannot be larger than the 2.5 nm par- ticles observed. Assuming, then, that the precipitates are 2.5 nm in diameter with a graphite crystal structure, we may estimate the diffusion path length for the carbon as follows: Graphite has a density of approximately Figure 7.3. 123 Log 10 90 (the time required to achieve equilib- rium) versus T/K (6 is the time required for the temperature to drOp one degree, r is the quench rate and (xc)Sat has been defined by Equation (7.9). The intersection of the hori- zontal lines with the log10 (90) versus T curve is the temperature below which, with the quench rate indicated, equilibrium cannot be maintain- ed by diffusion, e.g., at r 167 K'sec'l dif- fusion can keep the system at equilibrium down to 535°C and at r = 16.7 K-sec'l down to 950°C. 129 ORNL DWG 77-18725 ‘5 8=006msec /(X6)Sot'0.0053 (r-16.7OOK sec" ) -4 I— 3: o 6 m sec (mm-0.0026 (781670 K sec") -3 - 55 S O _I -2 I- 8860 m sec (r-16.7 Ksec") lIXCISat- 000062 -1 - (XCISat - B-GOOmsedr-IST Ksec") [0.00025 ' I l \ 800 600 - 400 TEMPERATURE/°C 125 2 ug/cm"3 or an atom density of 100 atoms nm'3. Nickel has a density of 90 atoms nm’3. A 2.5 nm diameter sphere has a volume of 8 nm3 and contains 800 atoms of carbon. If the carbon concentration is 0.0073 atom fraction (0.15 wt %), a volume containing 800 carbon atoms would contain 1.1 x 105 nickel atoms. A sphere containing 1.1 x 105 nickel atoms has a radius of 6.6 nm. The precipitates are taken to be at the center of spheres 20 nm in diameter. The diffusion path length is then 10 nm. The horizontal lines in Figure 7.3 are the time inter- vals required for the temperature to fall by one degree at various cooling rates. If for some temperature he < 6 (6 is the time required for the temperature to drop one degree), equilibrium is maintained and carbon precipitates to the extent dictated by its solubility in nickel at that temperature. When Me > 6, solubility equilibrium cannot be attained by diffusion. Carbon continues to precipitate, but slower and slower since the temperature continues its rapid decline. An independent estimate of the temperature below which precipitation ceases is obtained from the experimental result that the atom fraction of carbon remaining in solu- -h tion is 8.3 x 10 (0.017 wt %). The solubility of graphite is given by Equation (6.1), 126 log10 (xc)sat = 0.260 - 2816/T - (7.3) According to Spear and Leitnaker (1969), graphitic carbon which forms at temperatures below about 2000 K has a Gibbs free energy approximately 2.1 kJ mol'1 greater than true graphite. To account for this fact we add 2lOO/R J-k'1 to the enthalpy in term in the solubility equation. Equa- tion (7.3) thus modified reads: loglo (x0)sat = 0.260 - 2563/T (7.u) u The temperature corresponding to x0 = 8.3 x 10- is 756 K. At this temperature, he is 20 nsec and is rising rapidly. 1, would be re- A slow quench rate, less than 50 deg sec- quired for equilibrium to be maintained at this temperature. Until the time when He exceeds 6, (i;e;, at temperatures above about 800 K), diffusion is sufficiently rapid that equilibrium is maintained. 3. Previous Results Previously, Shriver and Wuttig (1972), Ulitchny and Gibala (1973), and Stover and Wulff (1959) have used optical metallography to infer that no precipitation has occurred in their quenched specimens. Our results indicate, however, that neither optical metallography at 127 1000x nor bright field TEM at 175,000x provides positive evidence that precipitation has not taken place; neither technique is always adequate. Shriver and Wuttig (1972) have measured the magnetic disaccommodation amplitude (the difference between the magnetic permeability preceding and immediately following demagnetization) of a Ni-0.3 wt % C Alloy. The magnetic disaccommodation amplitude is, according to Shriver and Wuttig (1972), proportional to the square of the amount of carbon in solid solution. This implies that the ampli- tude should continue to increase until all of the carbon is in solution. Their Figure 2 shows no change after 550°C; this indicates that the amount of carbon in solution was not changed by anneals at temperatures above 550°C. Equa- tion 6.1 indicates that 0.3 wt 1 carbon is not completely soluble until approximately 1070°C. After annealing at temperatures exceeding 550°C, the carbon in specimens of Shriver and Wuttig (1972) must have precipitated on cool- ing to approximately the equilibrium level at 550°C. Although Wuttig (1977) admits that precipitation occurred in his samples prior to the magnetic measurements he assumes it occurred at the annealing temperature. Since nickel carbide is not stable at the annealing temperature (Hansen and Anderko, 1958) and since carbon has been showntx> obey Henry's Law to the solubility limit in nickel, the possibility of the formation of a precipitate which would 128 lower the solubility of carbon to that at 550°C seems remote. If carbon were precipitating at the annealing temperature, the alloys would not reach equilibrium with graphite until all of the metal for the hypothetical carbide had been used up or all of the graphite had been transformed to the precipitate phase with the lower carbon activity. Ulitchny and Gibala (1973) measured the internal friction of several iron-nickel-carbon austenitic alloys. Internal friction peaks in austenitic alloys "have their origin in the stress induced reorientation of inter- stitial solutes which are paired (or clustered in larger numbers) with other point defects", (Ultichny and Gibala, 1973). Large changes are observed in internal friction peak heights as a function of quenching temperature and quenching rate. If the carbon clusters responsible for the peaks were the same as the residue we extract from nickel alloys, quenching temperature and rate would not affect the peak heights. Ultichny and Gibalas (1973) specimens contained 2 atom percent carbon. From Smith's results (1960) the solubility of carbon in iron-36 at % nickel alloys at 1000°C is 1.75 at z and by extrapolation is 1.15 at % at 900°C. Thus, all of the carbon was not in solution at two out of three of Ulitchny and Gibala's experimental temperatures. When the correction for the amount of carbon in solution before the quench is made, 129 the peak height per atom percent carbon in solution be- comes approximately independent of temperature, in agree- ment with our results. According to Ulitchny and Gibala the peak height is decreased by a factor of approximately 5 on slowing the 1 to 0.017 K sec-1. Now, the quench rate from 170 K sec- peak height is proportional to the number of carbon clusters and not to the number of carbon atoms in solution. By optical microscopy Ultichny and Gibala observed graphite precipitates in the slowly quenched specimens. Since the size of the precipitates increases during the slow quench, the number of precipitates decreases and the lower peak height results. The results of Ulitchny and Gibala (1973) are thus consistent with our both in terms of temperature dependence and quench rate dependence. E. Summary The fact that a carbon residue can be electrolytically extracted from nickel and nickel-titanium alloys contain- ing carbon has been established. The most likely explana- tion for the residue is that the carbon is precipitating during the quench in a first step in the dissolution of the super-saturated solution. This interpretation is consistent with the results of Shriver and Wuttig (1972) and of Ulitchny and Gibala (1973). The carbon "clusters" 130 that these sets of investigators discuss are very likely the residue that we have extracted. One consequence of the precipitation of free carbon is that analysis of electrolytically extracted carbides for carbon is considerably more difficult since carbon is present in two different phases. CHAPTER VIII THE NICKEL-TITANIUM-CARBON SYSTEM A. Results of the Carburization Experiments The results of the carburization of two nickel-titanium solid solutions are summarized in Tables 8.1 and 8.2 and displayed in Figures 8.1, 8.2, and 8.3. The addition of titanium to nickel increased the concentration of carbon, relative to that in pure nickel, at all temperatures studied (Table 8.1). At 1215°C, 2." atom percent titanium increases the equilibrium carbon concentration by 3.0%, at 1100°C by 9.0%, and at 900°C by 7.9%. Increasing the titanium concentration by 50%, to 3.6 atom percent, ap- proximately doubles the increase in the carbon concentra- tion. These results agree in magnitude and sign with the only literature values, those of Golovanenko 93 al., (1973). They reported the percent change in the concentration of carbon relative to pure nickel at 800, 1000 and 1200°C in an alloy containing 3.4 atom percent titanium and found, according to a plot in their paper, that the carbon con- centration was increased 18% at 1200 and 800°C and by 10% at 1000°C. They did only one experiment at each tempera- ture and used only one composition, so that uncertainty 131 1.322 Table 8.1 Experimental Results of the Carburization of Nickel-Titanium Solutions. Composition N1 N1 + 2.u at 1 T1 N1 + 3.6 at 1 Ti Data Percent Percent Set Temp./°C C, at 18 C, at ’a Increase C, at $3 Increase A-7783-nu 900 0.uu9 0.u68 u.2 0.098 10.9 A-7783-u5 0.256 0.285 11.1 0.292 10.0 A-7783-h7 0.201 0.215 7.0 0.238 18.3 A-7783-136b 0.211 0.230 9.1 Avg-7.9(l.5)c Avg-lh.u(2.1)c A-7783-u 1100 0.177 0.198 11.9 A-7783-17 1.05 1.10 8.6 A-7783-18 0.869 0.9111 8.3 1.03 18.5 A-7783-19 0.35“ 0.u08 15.2 A-7783-20 0.108 0.123 13.9 A-7783-35 0.661 0.825 2L.8 A-7783—32 0.816 0.909 16.3 A-7783—125b 0.303 0.32u 7.0 0.3u8 1u.8 Avg-9.0(1.0)c Avg-17.3(1.5)c A—7603—97 1215 0.637 0.653 2.6 A-7603—118 0.161 0.16u 1.8 0.172 6.8 A-7603-12l 0.211 0.215 1.9 0.215 1.9 A-7603-123 0.178 1.85 1.0 2.07 16.2° A-7783-ll6 0.332 0.350 5.u 0.366 10.3 A-7783-120 0.676 0.687 1.6 0.710 5.0 A-7783-123° 0.618 0.639 3.u 0.666 7.8. Avg-3.0(O.S)° Avg-6.u(1.h)° aConcentrations are relative to NBS SRM 195. b d calculation of the average. Equilibrium achieved by decarburization. cParenthesized uncertainties are apI-Ia/lfi where a I is the root mean square residual. Precipitation of Tic may have oCcurred in this specimen. The result was not used in the 133 Table 8.2. Activity Coefficient3 of Carbon in Nickel— Titanium-Carbon Solutions. 900°C 1100°C 1215°C Composition 0 b c «a b xa c (at %) Yo n CY Yc n oY Yc nb CY N1 136 8 1.” 6U.3 11 0.8 “2.0 7 0.8 7261 128 u 1.5 61.0 5 1.5 M1.1 7 0.1 N1+2.u Ti 7068 120 3 1.5 5u.0 6 0.9 39.3 6 0.7 Ni+3.6 Ti aActivity Coefficient calculated from carburization data and Equation (3.6). b Number of measurements. C0Y = 523 where c is the root mean square residual. /n Figure 8.1. 13“ Activity coefficient of carbon in nickel— titanium alloys at 900°C.. Ni + 2.11 at. 7 Ti; ONi + 3.6 at. % Ti; top line from Figure 6.1. Note that experimental error is exaggerated in that ye rather than 2n 7c, is plotted. A 7c, CARBON ACTIVITY COEFFICIENT 135 ORNL-DWG 77-9401R I I I I I45 I_ 4 ._1 900°C I35 Ni _ . . Ni '7' 2.4 '70 TI 125 F- _.1 . . o n _ _ _ 1133713— "5 -— O — I L . L I 0 0.25 0.50 0.75 1.00 ac, CARBON ACTIVITY 136 o o , .smssOHa mH w Cu can» eczema w page :a ompmpmwwmxm ma prcmefipmaxc own» 8.62 .Ha 83E 6c: as» .3. a .pm m.m + :8 .3. a... .66 :.m + H2. .ooooHH pm whoaam Eswcmpfip waofic CH QoDLMo mo osmfiofimmmoo zpfi>fipo¢ .m.w mpswfim 137 zommqo .10 E254 .66 00.. 050 00.0 0N0 _ _ a _ 0 .hm‘omfiiz o o o I IIIIIIIIIO . 0 O _._.o\u¢.N+_z I _z e 0600: m00¢mlhh 030....zmo 00 mm iN3IOIJJ300 Ail/“10V Noesvo‘ °é 138 .empsofia ma .6» as can» honpmh w pmcp Ca consummmmxm ma posse Hmpcoafiuoaxo pan» 0902 4.0 madman on: do» .3. R .pm :.m + 36.3. m .pm m.m + H2. .OomHNH um WNAOHHM. ESHCGPfip HmXOHC CH COQEMO .HO #COH0fihhmoo hpfi>fipo< no .m.m mpsmfim 2033 .._o >:>_»o< .66 00.. 0.1.0 00.0 0N0 _._.m.m+_z O/ QJ l'l'l'| _ Oomvmv mmmnmlbh 030......2m0 on A v 0 11130133300 All/\llov NOSUVO‘ 1M0 and composition dependence are unknown. Figures 8.1, 8.2, and 8.3 show the scatter, approxi- mately 3% at 900, 5% at 1100°C and 3% at 1215°C, and they show further that the carbon activity coefficient can be taken as independent of the carbon concentration over the ranges investigated. The decrease in the carbon activity coefficient (Table 8.2) upon the addition of titanium to nickel results in an increased solubility of graphite in the solid solution A-l sat = Yc ). Above a certain level of titan- (because (xc) ium, precipitation of titanium carbide occurs in nickel- titanium-carbon systems (Stover and Wulff, 1959). When the activity of titanium is large enough, titanium carbide can exist in equilibrium with both the nickel solution and graphite. Addition of more titanium to the system at this tricritical point at the same time decreases the value of the carbon activity coefficient and decreases the solubility of carbon in the solution. Table 8.3 contains the values of ARE, ASE and the parameters describing the temperature dependence of in ?c in the nickel-titanium-carbon solutions studied. From the results in Table 8.3 the composition dependence of in ?c could be fit with an equation of the type -<> II in in YC(N1) yc(Ti) 2n Yc(Ni) + 1n yc(Ti) lhl .deSUHmmh mhwddm 2008 0.00% 09.0 mepEHmpLQOCD UmNHWOSPmHmmAO .mHo.o 705 m + a. AN.HV m.Hm Am.mv m.mm As.mv o.am H-Hos.ss\mma Am.HV m.= AN.HV H.s As.mv o.m Huaos.s.aux\MM< “memo Home Aosmv News Aoozv omsm Ama.ov Hm.o- nom.ov ma.ou sham.v 06.0- nae R 0a m.m+az gas a as s.m+az paz o Iamefiz sfi m< can .mCOHpSHom conhwolssfiQMpfiB o o m< mo mmsfim> on» ocm mw mo monoccogom manpmthEmB one .m.w wanna 1&2 where in 70(T1) = c-xT1 + DaxT1 T'l However, the composition range studied so far is too small to warrant such a fit. B. The Solution Thermodynamics of Titanium in Nickel- Titanium—Carbon Solid Solutions Stover and Wulff (1959) made a careful phase diagram of the nickel-rich corner of the nickel-titanium-carbon system. When their data are combined with titanium carbide data from the JANAF Thermochemical Tables (1971) the ac- tivity coefficient of titanium at the graphite, titanium carbide, nickel solid solution tricritical point can be calculated, as follows: The equilibrium constant Kf for the formation reaction Ti(s) + C (graphite) = TiC(s) is the same as the equilibrium constant for Ti(in N1) + C(graphite) = TiC(s). Thus, for the three phase equilibrium here, 1H3 since A (graphite) = 1 = ATiC' (An additional point noted by Stover and Wulff (1959) and confirmed in this study (see Chapter IX) is the minus- cule solubility of nickel in titanium carbide. The low solubility of nickel in the carbide Justifies the assump- tion that the activity of titanium carbide can-be set to unity.) Table 8.“ contains the resulting activity coefficient values. One notes immediately that the partial molar excess free energy of titanium is large and negative. To calculate the partial molar excess entropyaand enthalpy a temperature dependent regular solution model is assumed. The values of the regular solution parameter, A, in Table 8.3 allow the excess functions at XTi = 0 to be calculated. 1 2n 7T1 = -20.6 T- -2.5, = 0.001 0 £nyTi E 1 K‘l, as = 0.8 J-mol‘ 1 -1 Ti K As = -21 J-mo1‘ E 1 T1 AH = -171 kJ'mol-l, 0H = 1.3 kJ-mol’ The assumptions in these calculations are that (l) nickel and titanium behave like a regular solution over the range 1AA 0 a .Uopomawoc soon was cognac mo CO005000pzoo one . mxflev< u Hoc< 000008 co0usaom amaswon pcoucoamu 005000006090 K 0 .0-0000000 - 090 0o00 00000000000 .0“. Aa.ovu .09x0AH.oVH 00chppoosz mumsfixopaa< .Ammmav «0053 cam po>opm Song copr mosHm>n .AHNmHV mmHDMB HacfiEmnothwSB m0 om 000 00 mpuua-mp toucfiscano ucmuo000000 uu0>uuo00>0po o r o > 0 vs 0: pm on o: a» no on aw 00000< 000000 0o0000 06000 .mCOHusaom 0000m 0100102109102 :0 :00» 10moudoo UCm 00000000509 no c00uousm 0 mm conpmo no pcm0o0mmmoo >00>0uo< .0.m magma 1N8 .00000000020 00 000 0000300 0000a 0 000000 .000HH0 01HBIHz 000 01Hz Ca 000000 00 30A 0.0000m 0000 HHH> 000 H> 00000000 :0 000H000H00fi 000 00 uanH 0H .000H3000 000 0000 0002 .003H0> 0w000>0 000000000 000HH 009 .ooOOHH 00 mmoaa0 E:H8000018500000H051ESH000fl01H030H0 00 000000 00 000H0000000 00H>H00< .H.m 005000 1119 202.3 .3 $5.84 an 00.. 0N.O 00.0 0N.O _ _ _ _ nmNh fl 4 anh ‘ ‘ CNN! _2 wab J! 500.QI_F¢.NI_Z 0 O Nwmh 606:5sz 4 oxméuttuni o 2286;5sz o 0.00: _ _ cnmnalhb OBOIJZCO 00 cm ON 00 malomaoo umuov mean “'1 V 150 .manmfimfiumsh ma pfig mmpmsvm pmmmH m maommn .mmoaam QIHBIHZ 0cm Ulfiz CH ommmno ma 3mg m.zpnmm was» HHH> cam H> mpmpmmzo ca macapmoficcfi map mo unwfia ca .ommfidcmp mam moms who: .mm:am> mwmpm>m pcmmmpamp mmcfia mag .OomHmH um whoaam ESHEOAQousscmcanoEI53ficmpHplaoxofic CH coppmo no pcmfiofimmmoo mufi>fiuo¢ .m .m mpsmfim 151 zommqo no 3.284 so 050 00.0 0N6 O _ _ _ OZN.m+_._.m.N+_Z O osm.¢+;¢.~+_z 0 TI .00.m+_._.¢.N+_Z d II. on bow.¢..._._.n.m+_z d 99$ ‘ 4 Q 4 nwmh a4. d d H mwmfi d 4 1 II 4 O¢ 4 _2 o o o vows. o o o I I. on New» 0 o. _ _ _ momnmlhh 030 IJZKO V O malmssaoo umuov ‘ 152 The narrowness of the solid solution region increases the difficulty of the carburization experiments. In particu- lar, alloy 7266, which contains the largest concentra- tions of both molybdenum and chromium has a single phase region so narrow that quantitative data on the solution phase were not obtained from carburization experiments. Instead, annealing experiments discussed in Section 8.2 were performed in order to obtain data on this limited region. Although alloys 7267, 7268 and 7262 also have small carbon solubilities, it was possible to obtain quan- titative carburization data on all three solutions at 1100 and 1215°C. To determine effects of alloying additions on the activity coefficient of carbon two procedures can be followed: (1) compare the activity coefficients of carbon as determined with the iron standard equation (3.6); or (2) compare directly the difference in carbon concentration of two alloys in equilibrium with the same gas composition. The second method is necessary for some of this work be- cause not all of the alloys were present in every run and therefore the effect of the iron standard does not cancel out. Such comparisons are shown in Table 9.2. Compared to nickel + 2.H titanium, molybdenum decreases the equilibrium concentration of carbon from 12% to 19% at the H atom percent level and from 15% to 25% at the 8 atom percent level (Table 9.2). Percentage increases 153 H.H m m.m ea.m efi.m- + mm mum wmmw >.H m o.m 02.: so.: o.m a.a+ 0: 0.: WWW» m.m m NH- o.m m 3H- om.m H emfiu 0: m.: WWW» m.m m ms- mo.m H sow- m:.m emm- 0: N.w .Wmmw m.H m CH H.H m CH m.o m m.H so =.m mmms m.H m o.m o.m : H.m m.H : o.mu to m.: mmmw N.H m ma m.m 2 3H o.m m 0.0 so o.m mmmw we on owcmco we a: mwcmno we as mmcmso coapfimoqeoo nfimm mucoopom mucmohmm apcoopmm “hmowcmno moHH< oomama ooooaa Oooom .mCOfip usaom o-ponozufiaufiz ea copgmo mo mcoapwppcmocoo Esfipnaafiscm mo comHAmQEoo .m.m magma 154 .sz>Hpom mEmm on» no mcodeHOm 03» map so muse a H m N» H m» p mo xowH m on one COHmem chp :H popmHSOHmo mm; o .mnm>\ rv. mo + Am>\Hv. mo u mom m . > .ooH x.lIII| u ownmso pcoopom .omm: ohms H.m cam H>IN> m.m mmHnt CH mpcmHonmmoo sz>Huom map comewano pomnHo you womb mo xomH m on once .Hmsvamn chandm cmmE poon map mH b when; m\\o u poo .mpcmEmpSmmmE mo noneazn m .HoHHm ommeHeeH on» :H GOHpmppCmozoo congmo on» mH Ho mama: .00H x IILWI u mmoHH< "owcmno accommm wonHo HNOHH¢ m 0 O O O . a“ 0 NF H H m s: 0H N am mm . o a m meme no a.m . . . . o . Hmma m o m 0 mm we s cm :H + s H H meme mofi ®.© .®._.~ H.m| . 02 N.@ m NF m U Nmmn we as mwcmno we a: owcmno we a: owcwno QOHpHmoqeoo gHmm muzmopmm mucoopmm wpcmopmm CH mwcmno mOHH< oomHmH OOOOHH . cooom .omscheoo .m.m «Hams 155 are larger at lower temperatures. No literature exists on the effect of molybdenum on the equilibrium carbon concentration in nickel solutions. The value for the Kohler-Kaufman parameter (wMoC) estimated by Kaufman and Nesor (1975) indicates that molybdenum should decrease the equilibrium concentration of carbon in nickel solutions, as found here. Wada £2.2l- (1972) indicate that molyb- denum increases the equilibrium concentration of carbon in iron solutions, opposite to the effect on nickel solu- tions. Compared to alloy 7261, chromium increases the equilibrium concentration of carbon in nickel at 1100 and 1215°C but has no effect at 900°C (Table 9.2). The decrease is from 3% to 6% at the u.6 % level and from 14% to 15% at the 8.0 % level. Golovenenko gt a1. (1973) measured the equilibrium concentration of carbon, relative to nickel, in a solution containing “.0 at % chromium at 800, 1000 and 1200°C. They found that chromium decreased the equilibrium concentration of carbon by 15% at 800°C, 6% at 1000°C and 3% at 1200°C. Neither the temperature dependence nor the sign of the effect of chromium on the equilibrium concentration agrees with our results. Golovenenko gt a1. (1973) did not estimate the size of their errors. Chipman and Brushy (1968) reviewed the data on the effect of chromium in iron and indicate that 8 atom percent chromium increases the equilibrium concentration 156 of carbon in nickel by about 7%. The reason for this large difference is discussed in Chapter X. In the more complex solutions containing both chromium and molybdenum, the effect of additions on the equilibrium concentration of carbon is more complicated. The addition of 8 at.% chromium to a solution containing u at.% molyb- denum (726M + 8 at.% Cr + 7268) increases the equilibrium carbon concentration by as much as ”5% (Table 9.2). From the previous discussion one would expect the carbon concen— tration to be increased by W15%. Similarly the addition of 8 at.% molybdenum to a solution containing u atom percent chromium (7265 + 8 at.% Mo + 7267) has little effect at 1100°C and increases the equilibrium concentration of carbon by 6.8% at 1215°C. The results for the addition of 8 at. % molybdenum to alloy 7261 suggest that the equilibrium concentration should be decreased by from 15% to 20% upon the addition of 8 at.% molybdenum. The relative change in the equilibrium concentration of carbon depends on the amount of both molybdenum and chromium added (Table 9.2). In the case of chromium a much bigger relative change’ takes place upon the addition of 8 at. z than “.6 at. %. The addition of u at. % molybdenum on the other hand has larger relative effect than the addition of 8 at. %. 157 B. Carbide Precipitates The solubility of carbon in equilibrium with the metal carbide that forms in these alloys was determined in two different ways. In one set of experiments alloys of fixed composition were annealed at the desired temperature and then quenched. The amount of carbon in solution was determined from knowledge of the bulk carbon concentration, the weight percent of precipitate in the alloy and the concentration of carbon in the precipitated phase. This method is particularly suited to alloys with low carbon solubility. In the second method, the solubility of car- bon was determined from the break in the concentration versus activity curve obtained from gas phase carburiza- tion experiments. The concentration above which the atom percent carbon in the alloy is no longer directly propor- tional to the activity of the carbon is the solubility limit. This method is better suited for alloys of high carbon solubility. l. Carbide Composition The precipitates extracted from the carburized alloys were analyzed with an electron microprobe, by the method described in Chapter V. Table 9.3 contains the results of these analyses together with the lattice parameter 158 .3 0532008 0:032. 05 8303588 8008 HHS 2: .3 0003.65. 0 sec 000E300 83.0 .00050 Us 0Hnso 0:» 0:0 wanna umHuHucooch an up omunH0000 000:0 oanuao one: .0m com on cow 3000 youoaopoauuqu naHh cognac: hHHuauo: can: naoaHoonm .aneHuou vacuum aqua «ooh as» 0H 0 cheap I: Hooo.Ouou .0oHpcuaaoun noun. an vonHappo on. 0 cheap -.o-0 .HasuHuvu 090500 0003 0009 «:0 0H 0 onus. nooHuo .hhaoaouuooau 00¢: nausea Joann hp coca-00009 Hzo .H090Huou oudsdu :00: finch 0:» 0H 0 090:. .uno.000 .coanuoona 0Haovu hp voaHauovoo no u v .09” I o u .ub o “H.783 «.0: 00H0000200000 canyon on» yo muvloocx sou» voaHaapo au=H¢> ovaHouA< .ononnouoHa couauoHo osa saHh voaHauouoc ooHaau .uaubo 30:38.3 «00200 20H? 30 05 03H. 05 5 8008 0o 0:032 2: .031 23 0H 03.30 .06Hegwaoun.:xflo_buuoHL0 n .HdauHuou oudsao ance «och on» 0H 0 cheat n .pl «Ho.onoa u .04 a H30 05.0 00.0 00.0 0.0 0H 0H N0.H 00HH 070204 .5 0.0. 00.0 N00 00 0H 0H 00.N 00HH 3.00004 2. 0.?! 33.0 0H.H 0m .0 00 0H 0H 00 .0 0HNH 00-00004 00N0 300.0 00.0 00.0 N0 NN 0H 00.... 8: 0100004 0 :2 00.0 H0.0 mm 0N NH 00.0 8: 3-00004 0: Nd. 00.0 00.0 00 HN NH 000.0 8: 3.3004 .8 0.? S? .0 8H 00.0 0N NN 0H 3H 0HNH 00.00004 8. “NH: 000....0 N00 0. H 0N 0N NH :34 0HNH 00.00004 00N0 . - .- u .04 0N2 0 00.0 H00 H0 0H NH 000.0 8: 00-00004 0: N4. 300.0 00.0 00.0 00 0H 0H «AN 8: 0700004 H... 0.?! 0N0. .. . 0 N0.0 N0 . 0 00 HN 0H 00.N 0HNH 02-00004 3N0 N00 00 . 0 0N 0N .HH .0 00.0 82 00.00004 00.0 00H 0N N0 HH 000.0 8: 3.3004 n .0. 300.0 00.0 00.0 00 0N HH 00H 82 0.00004 .8 0.0 80....0 00.0 00.0 0N «N «H N0.H HHNH 00-00004 .0: H0. 003.0 N00 0.0 00H 0N 00 HH ”8.0 22 00-00004 H0 «NH: 00 .0 0 .0 0.N m0 . H 0N 00 NH c00.0 0HNH 00.00004 00N0 200.0 00.0 00.0 0N 0N NH 30.0 8: 0700004 0 .00 00.0 0....0 0N 0N NH .HHN 00: 3-00004 0: N0. 20.4.0 N0.H 00.0 0N NN 0H 0H.N 0HNH 00-00004 H0 0.N.Hz 200.0 00.0 00.9 00.9 00 .0 00 ...N 0H 00 . H 0HNH 00-00004 N0N0 0&2 0300 Hz .00 .5 00 H0 H0. 00 o: .00 3030320 .HoHH4 0.: 5 H000 000.52 00H: oanuoo 0o uHuaozuu mu aoe< cu aca< II. on aoa< on acu< 0:0 :H opouHQHoopm .9909 oHnadm 0383.300 0: no u.‘ cu.‘ 000H0000 .coHuomnuhHo pennx cam anLMOuoH: comuome on» an mouMuHaHovhm vanhwo on» yo uHumec< on» go nuHsmop one .m.m anme 159 of the precipitate phase as determined by powder x-ray diffraction. The weight percent carbon in the precipitate phase was calculated through a knowledge of the bulk carbon concentrations, the weight percent precipitate, the activity of carbon in the specimens and the activity coefficient of carbon in the alloys. In this way the con- centration of carbon in solution is calculated directly and the concentration of carbon in the precipitate by difference. The method for calculating the molybdenum and titanium concentration is contained in Chapter V. From Table 9.3 it appears that the Mo/Ti ratio in the precipitate depends on the amount of molybdenum in the matrix. It also appears that the ratio increases as the weight percent precipitate in the alloys increases. The Mo/Ti atom ratio in the cubic precipitates formed in the allomscontaining u atom percent molybdenum (726“, 7268) is 0.H2 i 0.003. The value of 0.62 obtained for alloy 726“ A7603-l23 (Table 9.3) is inexplicably high. The lattice parameter of the 726“ A-7603-123 precipitate is not different from those of the other two 726R specimens, both of which have lower molybdenum concentrations.- Doubl- ing the molybdenum concentration in the matrix, to 8 at. z, increases the Mo/Ti atom ratio in the cubic precipitate phase by almost 100% to 0.79i0.0l. Nickel and chromium are minor elements in the 160 precipitate phase. Chromium is more soluble in the carbide than nickel, but it is likewise depleted in the precip- itate phase relative to the matrix. Figure 9.3 shows the effect of changing the molyb— denum concentration in the carbide on its lattice param- eter. Over the range explored (Mo/Ti atom ratio 0.“ to 1.0), the lattice parameter is a linear function of the MozTi ratio in the precipitate. The addition of molyb— denum decreases the lattice parameter of the carbide. Alloys 7268 and 7267 differ from alloys 7268 and 7262, respectively, only in that they contain 8 at. % more chrom— ium in the matrix. The addition of the 8 at. % chromium to the matrix lowers the precipitate lattice parameter by approximately 0.0005 nm. The effect of chromium on a per atom percent basis is larger than that of molybdenum, presumably because of chromium's smaller atomic radius (Slater, 196“). As shown in Table 9.3 the carbon-to-metal atom ratio in the precipitate was almost always less than 1. The average value is 0.85, o=0.ll, and o//fi=0.03. The Ti-Mo carbide might be viewed as a solid solution between nearly stoichiometric TiC and Mo3C2. Molybdenum increases the lattice parameter of nickel at a faster rate than does titanium, yet molybdenum is observed to decrease the lattice parameter of TiC. Since the lattice parameter l6l .m.m mHnme cH 0000 on» 00 0H0 0000000 pmmmH m an UmcHEpmpmo was mcHH one .mcanmo on» :H HB\oz mo COHpocsm 0 mm mumpHQHoopQ mchpmo 0:» mo popmsmpma moHpumq ”0.0 mpsmHm 162 C. 0.9.3 02 0.40:5 0: N.0-H0 0.N-Hz / f . .0: N.H-H.H 0..N..Hz .000 .. 223.5012 .0 0.0..02N.0..:0.N-_z 50.0.02 .0 130.2 ouc¢>< mmnmuhh OBOIJZGO Onvd .nvd vad nmvd (W) 00 aouaavo 161 .m.m mHnme CH apmo on» mo pHm mmpmsvm ammoH m an UmcHEpmpmv 003 ocHH one .oanpmo on» cH HB\oz mo COHuocsm 0 mm mumpHQHompa mothwo 0:» mo nmuoEMAMQ moprwq 00.0 mpstm 162 .0002! 22 000:5 OS N.mIH.H. m.NIHz .02 m.zIHB :4..le covdlothtfimdlz .UV.¢I02N.mI_Pm.NLz com.mIo2_..m|_._.¢.NLz o-ocx mmnmuhb DBOIJZmO Onvd .mvd mnvd nmvd (um) 0:2 acneavo 163 of Mo is 0.U28 nm and that of TiC is 0.N33 nm, a ready 302 explanation is provided by a TiC-M02C3 solid solution for both the lowering of the carbon lattice parameter by molybdenum and the substoichiometry. 2. Annealing Experiments Table 9.“ and Figure 9.h contain the results of the annealing experiments. Since the weight percent precipi- tate extracted from alloy B (Ni + 2.1 at. % Ti) did not change as a function of temperature, we infer, with the help of the evidence of Chapter VII, that the extracted material precipitated on cooling. This means that at least 0.08 wt % carbon is soluble, in alloy B, at all the temperatures investigated. Alloy C (N1 + 2.” at. % Ti + 8.2 at. % Cr + 0.5 at. % C) behaves like alloy B at high temperatures. The weight percent precipitate extracted from alloy C annealed at 1100°C is equal to that from specimens annealed at 1260°C. At 760°C, however, the weight percent precipi- tate increases by a factor of two. The solubility of carbon in alloy C at 760°C was calculated on the assumption that the precipitate was stoichiometric TiC and that 0.07 wt % of the precipitate formed during cooling (see Chapter VII). The value of 0.0fl5 wt. 1 for the carbon solubility at 760°C should be considered a minimum estimate since 1614 .lifv ¢.-. Fesults .f the Annetlinr Fxrerirents. ' wt. Sc wt. 5‘ wt. Sb Carbon Carbon in Precipitate in Solid 421‘“ Ste‘ireu Annealing Annealed in Annealed Solution 3‘. ' hunter Ristcrv Tcmp./°C Time/hr specimen specimen Solubility a x1+2.‘ *1 BIS As received 1260 16 0.08 0.1111d 0.08 815A 1: hr 3' :fr1 760 168 0.08 0.157d 0.08 N1+2.n Ti C-6 As recuizvd 1200 16 0.103 0.119d 0.1 +8.: Cr c-7 As 00:01:»1 1200 16 0.098 0.117d 0.1 C-6-A :6 hr a' 1210 760 168 0.108 0.359 o.ou5° A 321 As received 1260 . a 0.083 0.037 0.078! Ni+2.6 Ti A-B As received 1260 16 0.102 0.093 0.090 +8.“ Ho A-lO As received 1260 16 0.092 0.037 0.087 BAIH As received 1200 1 0.102 0.10“ 0.088 BAZH As received 1200 2 0.102 0.095 0.090 A-7783-1u7 u hrs at 116C 1100 18 0.083 0.328 o.ouo A-7753-5 ' As received 1000 72 0.078 0.h21 0.023 A-7783-5 As received 900 118 0.083 0.533 0.013 A—7783-5 As received 800 500 0.08h 0.626 0.002 A-B-A 16 hr at 1‘60 760 100 0.096 0.732 ----- “99 1177 2 0.035 0.050 0.028 N1+2.5 71 A-7783-1u7 “ hrs at 11(0 1100 18 0.0265 0.070 0.017 07.2 no A-7783-5 ‘8 received 1000 72 0.0275 0.126 0.011 + 8.8 Cr A-7783—5 A: rcceived 900 111 0.0303 0.223 0.001 A-7783-5 As received 800 500 0.0306 0.233 ----- 760 100 0.035 ‘ 0.28 ----- a a - 3! where a is the root mean square residual. be - 0.015 wt. S where o is the root mean square residual. cThe solubility of carbon in alloys hi9 and A was calculated on the assumption that the weitht percent carbon in the precipitate was 131. This was based on the assurption that all of the carbon in the stnctrurs 35x98}- ed at 760 had precipitated. Solubility - bulk carton concentratirn - (wt. 7 prt) x 0.11. The rvsul's frr the weight percent carbon in the precipitate found in Table 9.? indicate a value of «10' fer o. 6This precipitate was free carbon as described in Charter VII. eThe solubility of carbon in alloy C was calculated on the assumption that the precipitate wn: 3‘0 0 ichismetric TiC and that 0.0? wt. % precipitate resulted from the precipitation of free carbon (see Chapter .II). 1.The solubility for this specimen antears low. It may be that the precipitate was free carbcn. 165 Figure 9.“. (a) The concentration of carbon in alloy A (Ni + 2.6 at. % Ti + 8.“ at. % M0) in equi- librium with the cubic carbide phase as a function of temperature 0 = 10% of the bulk carbon concentration. (b) The concentration of carbon in alloy ““9 (Ni + 2.0 at. % Ti + 8.3 at. % Mo + 8.“ at. % Cr) in equilibrium with the cubic carbide phase as a function of temperature. a = 10% of the bulk carbon concentration. I00 [C '13-] 166 TPC) out-0's 774m INF-1'9] Ml “(caucus-#9 -u ._ _ JL 1 l 7 I .0 WT!!! cam-owe 77-13518 rm 4 0 8200 9000 300 ' l T I -2.0 — .— -3.0 —- — ALLOY 449 he peuq- 4.32 - 1:99 l J I I -Q.0 7 8 9 90.00%“) 10 167 TiC is often substoichiometric in carbon. Alloy A (N1 + 2.6 at. % Ti + 8.“ at. % Mo + 0.5 at. % C) has considerably smaller carbon solubility than either B or C. Figure 9.“a is a plot of the logarithm of the carbon solubility versus the reciprocal of absolute tem- perature. The solubility was determined on the assump— tion that the solubility of carbon at 760°C is zero and that the weight percent carbon in the precipitate is not a function of temperature. The addition of molybdenum lowers the solubility of carbon from something over 0.08 weight percent at 760°C in alloy B to something less than 0.001 weight percent in alloy A. Molybdenum lowers the carbon solubility relative to the carbide by three dif- ferent processes: (1) molybdenum dilutes the nickel- titanium solution and thus increases the titanium activity; (2) molybdenum forms a solid solution with TiC (see IX B.l) and the activity of the carbide is thus lowered; (3) the molybdenum-carbon interaction is weak relative to the nickel-carbon and titanium-carbon interactions, and the addition of molybdenum to the solution increases the carbon activity coefficient. All three of these effects tend to displace the reaction Ti(Ni) + C(Ni) 2 TiC(solid) to the right. 168 Alloy ““9 (Ni + 2.0 at. % Ti + 8.3 at. % Mo + 8.“ at. % Cr + 0.18 at. % C) results from the replacement of 8.“ at. % nickel with 8.“ at. % chromium in alloy A. Alloy ““9 and 7266 are essentially the same. Table 9.“ and Figure 9.“b show that the addition of the 8.“ at. % chrom- ium lowers the solubility of carbon relative to that in alloy A by a factor of approximately 3 at 1215°C. The decreased solubility of carbon in alloy ““9 is due pri- marily to diluting the nickel-titanium interaction which results in a higher titanium activity. That is, the Gibbs free energy of mixing for titanium and chromium is much less negative than for titanium and nickel. The chromium does not form an appreciable solid solution with the car- bide phase, and therefore the addition of chromium does not alter the activity of the carbides. 3. Carburization Experiments Table 9.5 and Figures 9.5, 9.6 and 9.7 contain the result of the gas carburization experiments undertaken to determine the solubility of carbon in various nickel alloys. Since it has been shown in Chapters VI, VII and IX that the carbon in solid solution in these alloys obeys Henry's Law, any negative deviation from Henry's Law can be considered evidence that carbide precipitation has taken place. The solubility limit is the concentration at which the 169 Table 9.5. Solubility of Carbon in Several Nickel-Based Alloys as Determined from Carburization Ex- periments. Temp. 900 1100 1215 (°c) Alloy gngi A:(sat) C(sat) A:(sat) C(sat) A:(sat) C(sat) wt.% wt.% wt.% 7262 0.17 0.019 0.18 0.0“6 0.18 0.073 726“ 0.36 0.0“5 0.38 0.10 0.32 0.1“ 7266b 0.0“6 0.016 0.067 0.037 7267 0.10 0.032 0.095 0.050 7268 0.11 0.0“3 0.13 0.079 7262 Ni + 2.5 Ti + 8.2 M0 726“ N1 + 2.“ Ti + “.2 Mo 7266 Ni + 2.“ Ti + 8.1 Mo + 8.3 Cr 7267 Ni + 2.5 Ti + 8.2 Mo i.“-" Cr 7268 Ni + 2.5 Ti 1 0.1 Mo + 8.“ Cr aThe solubility was determined from the following eguation = A . Activit coefficient 0 car on was AC(sat) Ye xC(sat) Y obtained from Tables 8.2 and 9.1. bThe activity coefficient of carbon in alloy 7266 was not experimentally determined therefore an approximate value had to be used. The activity coefficient of carbon in alloy 7266 was taken to be the average of those for alloys 7267 and 7268. This seems to be appropriate since alloy 7267 and “ at. % more Mo than 7268 and Mo and Cr have opposite effects. Figure 9.5. 170 Atom Z carbon versus activity of carbon in several nickel-based alloys at 1215°C. The intersection of the two lines, with the same label, is the solubility limit of carbon relative to the carbide phase. The lower line represents the solid solution where the slope is 100/9c (xC = AC/90). The dashed lines are an extrapolation of the solid solu- tion lines and represent the amount of carbon in solution at any given activity. The upper lines have been fit by least squares to the data from the two phase region, points that diverged from the straight line behavior exhibited near the inter- section were ignored. ATOM PERCENT CARBON 171 ORNL-DWG 77-9398 l | l l l l A o 1215°C , 7262 Ni-2.S Ti-8.2 Me (at. I) 726'! 111-2.0 71-0.2 lo (It. 5) 7266 Ni-2.“ Ti-8.1 Ho-8.3 Cr (at. S) 7267 211-25 T1-0.2 140-0.“ Cr (at. S) C) 7268 Ni—2.5 Ti-“.1 Ho-G.“ Cr (at. I) 7267 7266 e 7268 7264 . l l l 0.2 0.4 0.6 0.8 ac, ACTIVITY OF CARBON Figure 9.6. 172 Atom % carbon versus activity of carbon in. several nickel-based alloys at 1100°C. The intersection of the two lines, with the same label, is the solubility limit of carbon rela- tive to the carbide phase. The lower line represents the solid solution where the slope is 100%?C (xC = Ac/yc). The dashed lines are extrapolations of the solid solution lines and represent the amount of carbon in solid solu- tion at activities exceeding the solubility limit. The upper lines were fit by least squares to the data from the two phase region. ATOM PERCENT CARBON u N 173 ORNL-DWG 77-9396 l I l l I r T 7266 7267 7262 — —( F 1100°C “ 7262-Ni-2.5Ti-8.2M0 7267 - Ni - 2.5Ti - 4.4Cr - 8.2 M0 7268 - Ni - 2.5Ti - 6.4Cr -4.4 M0 7266-Ni-2.4Ti- 8.3Cr-8JMO _l / 7268 /7257 ’ 7262 I I O 0.4 0.6 ac, ACTIVITY OF CARBON Figure 9.7. 1?“ Atom % carbon versus activity of carbon in several nickel-based alloys at 900°C. The intersection of the two lines, with the same label, is the solubility limit of carbon rela— tive to the carbide phase. The lower line represents the solid solution where the slope is 100/?c (xc = Ac/90). The dashed lines are extrapolations of the solid solution lines and represent the amount of carbon in solid solu- tion at activities exceeding the solubility limit. The upper lines were fit by least squares to the data from the two phase region. ATOM PERCENT CARBON 4.25 0.75 0.50 0.25 175 ORNL-OWG 77-9397 . I l l l 7267 7264 7268 7262 Ali-270 I, 7264 / / I, 7262 A / / / /’ / — I/ / o ’ 900 C 7262-Ni- 2.5Ti-8.2 M0 7264- Ni- 2.4Ti- 4.2Mo 7268- Ni- 2.5Ti-8.4 Cr - 4.1Mo 7267-Ni- 2.5Ti- 4.4Cr - 8.2 MO I I l | 0.4 0.6 ac, ACTIVITY OF CARBON 176 concentration versus activity line for carbon in the alloy has a change in slope. In Figures 9.5, 9.6 and 9.7 the solubility limit has been determined by fitting the solid solution carburization data and the carburization data from the two phase region with least squares lines and calculating their intersection. The solubility of carbon in molybdenum-free alloys was not determined by this technique because either the carbide phase does not exist in the alloys at the tempera- tures and activities investigated or only one data point in the two phase region existed. The solubility limit of carbon in alloys 7266, 7267 and 7268 was not determined at 900°C because a diffusion barrier, possibly a layer of chromium oxide, slowed the rate of carburization so much that carburization experiments were impractical. As Table 9.5 indicates, doubling the molybdenum con- centration reduces the carbon solubility by a factor of 2. The result of adding chromium to the carbide forming alloys has a similar effect. Both the decrease in solubility of carbon upon addition of chromium and the values of the solubilities agree with results obtained for similar alloys in the annealing experiments discussed in the previous subsection. The effects of additions of chromium and molybdenum on the solubility of carbon relative to the carbide phase 177 in the alloys already forming a carbide phase thus follow a regular pattern: doubling the molybdenum or chromium concentration decreases the carbon solubility by a factor of about two. 0. An Unidentified Phase of High Carbon Content In alloys 7266 and 7267 some specimens contained an unidentified carbide phase (see Table 9.3). The Mo/Ti atom ratio is approximately 1.6 and the carbon to metal ratio in the two phase precipitate is approximately 0.8. Microprobe examination of precipitates, in the matrix (see Figure 9.8) revealed that the precipitates with the needle like morphology had the same Composition as the more rounded precipitates. The new phase does not correspond to any of the low carbon carbide such as M2C, M6C or M12C. Attempts to index the x-ray diffraction characteristic of the phase have failed as have attempts to identify it with the ASTM x-ray card file. Tables 9.6 and 9.7 contain the 26 values and relative intensities of the diffraction peaks in the spectrums for 7266 specimens A-7603-97 and A-7783-37. Figure 9.8 is an optical micrograph of the precipitates in alloy specimen 7266 A—7603-97: the needle- like morphology is not characteristic of TiC precipitates. Table 9.6. 178 X-ray Diffraction Data on the Unidentified Phase Alloy 7266 A-7783-973. 26 I 27.29 11 36.70 1100 “1.“8 85 ““.30 25 “6.35 20 51.11 130 5“.82 80 58.95 30 61.37 30 63.17 “8 67.67 15 72.“3 5“ 78.07 100 aCopper K0 radiation was used. speed was 1/“° 26 per min. The spectrometer travel 179. Table 9.7. X-ray Diffraction Data on the Unidentified Phase in Alloy 7266-A-7783-37.a 20 I 20 I 27.30 7 63.28 18 33.02 2 67.7“ “ 35.56 7 72.“5 22 36.70 870 72.65 1“ 37.0“ 73.73 6 39.00 “ 76.8“ “ “1.53 21 77.10 3 ““.“l 6 78.11 100 “6.37 9 78.33 51.12 2 88.20 2 51.29 1 88.35 2 59.00 9 88.“3 1 61.“5 6 90.38 2 aCopper K radiation was used. The spectrometer travel speed 1/fi° 26 per min. t 180 Figure 9.8. Specimen number 7266 A-7603-97 equilibrated at 1215°C at A0 = 0.268. Note the needle like precipitates which are characteristic of the unidentified phase. The other pre— cipitates are the MC phase. If 20 0.001 40 I 181 60 MICRONS 4— 500 ——lr INCHES 120I HO 0.005 CHAPTER X THE KOHLER-KAUFMAN EQUATION A. Calculation of the Nickel-Carbon and the Iron-Carbon Interaction Energies Table 10.1 contains the values of the “ interaction energies that describe the nickel-carbon and the iron-car- bon systems. The relative lattice stabilities are listed in Table 10.2. The equations used to calculate the inter- action are from Equation (2.30). For each temperature CE _ _—FCC- -gr+ 2 xN1(1'2xc) wNiC 2xCx Ni wCNi GE _'—FCC- -gr 2 2 2xCxFe wCFe' (10.1) To obtain wNiC and erC’ Equations (10.1) are solved at xC = 0, where A°° *FCC- RT 2n YC(N1) - G gr “N10 —FCC-gr 182 183 Table 10.1. Calculated Values of Nickel-Carbon and Iron- Carbon Interaction Energies, wigc.a -l -1 -l -l -2, -3 A13 kJmol Bid/Jmol K Cij/Jmol K 10 (113 b wNiC -135.52 87.31 -29.29 b WCNi ‘163o7 l“.0 C d erC - 96.15 -0.88 0.0 d wCFe -156.1 0.0 0.0 aw = A + B T + C T2. 13 id 13 1J b 1 °“=O'3 kJmol' , calculated assuming a 3% error in ?C(Ni). CAssumed zero. d 1 °“=O'2 kJmol' , calculated assuming a 2.5% error in ?C(Fe). 18“ Table 10.2. Some Relative Lattice Stabilitiesa for Elements of Interest. c 0 Element Transformationb 'H'E-a/kJ-molul 'E'Iji-a/J-molDJ'K-1 C Graphitic FCC 138 15 T1 BCC FCC -l.0 3.8 Cr BCC FCC 10.5 0.63 Fe FCC FCC 0 0 Ni FCC FCC 0 0 Mo BCC FCC 10.5 0.63 aFrom Kaufman and Nesor (1973, 1975), Uncertainties not stated. bFCC=Face Centered Cubic, BCC=Body Centered Cubic. c—b-a _ b —a b-a _ —b —a ' —b-a _ -b-a —b-a Hi - (21 - Hi), 6: - (s1 - Si) and G1 - Hi - Tsi . 185 To solve for wCNi and wCFe’ Equations (10.1) are evaluated at other values of xC. In the case of the nickel- carbon system, 7C is a constant to the saturation limit, and to insure that the interaction energies reflects this we evaluate wCNi at (xc)sat' For the iron-carbon system the results of Smith (19“6) in the form of Eq. (3.6) were used to determine the values of wCFe and erC from Eqs. (10.1) and (10.2). Experimental values for 7C and (x0) from Chapter VI were used together with EEOC-gr sat estimates of Kaufman and Nesor (1975) to obtain wNiC and wCNi at 900°C, 1100°C, and 1215°C. Data for nickel were fit with an equation of the type T2 (10.3) = A1d + B1JT + C1;] “’11 B. Analysis of the Nickel-Iron-Carbon System Smith (1960) and Wada 33 a1. (1971) studied the nickel- iron-carbon system from xFe = 0 to 1.00. Tables 10.3 and 10.“ contain the results of these two investigations. The appropriate Kohler-Kaufman equation for 6% at xC = 0 is E —FCC-gr Gc = RT in YC ‘ G0 + xNiniC + xFeerC (10°“) 2 2 ‘ xNixFewNiFe ' xFele‘VFeNl 186 Table 10.3. The Reanalyzed Results of Smith (1960) for the Activity Coefficient of Carbona in Nickel- Iron-Carbon Alloys. Mole Fraction Activity Coefficient rms' Nickel of Carbon Residual xN1 Yc O 0.0 8.“5 0.2 0.0379 10.5 1 0.0775 13.2 1 0.1“8 17.3 1 0.258 29 3 0.395 5“ 6 0.599 119 9 0.787 1“8 7 0.99“ 87.6 “.5 aTable contains values of 9: calculated for xC < 0.02. When xC XC+XN1 XC+XN1 - “0N1 ( , (10.6) 195 where ternary interaction energies are excluded. At x0 = 0, Equation (2.30) yields for 0%, E -FCC-gr G0 = ET in Y0 ‘ G0 + xN1¢N1C + xT1‘pT10 I - x2 x w - x2 x w (10 7) Ni Ti N1T1 T1 N1 TiNi’ ' where ternary interaction energies are excluded. Following Kaufman and Nesor (1975) we assume wTiC = wCTi and wNiTi = wTiNi‘ In both cases this is Justified because of narrow range of experimental data. These as- sumptions result in a symmetric excess Gibbs free energy as a function of composition in the binaries. While Eqs. (10.6) and (10.7) could in principle be solved simultan- eously they are easily solved by iteration. The estimate of wTiC (= wCTi) proposed by Kaufman and Nesor (1975) was used in Eq. (10.“) to solve for wTiNi (= wNiTi)' Then a value for wTiC was calculated from the results in Chapter 8 and Eq. (10.7). This value for wTiC was then used to recalculate by Eq. (10.6) the value of wNiTi' Kaufman and Nesorls (1975) estimate was close to our calculated value and only one iteration was necessary. The use of the value of wTiC obtained Eq. (10.7) to re- calculate wNiTi changed the value of wNiTi by approximately 0.“ KJ-mol'l. Recalculation of wTiC produced no significant change. The values for wNiTi at several temperatures were 196 fit by least squares to Eq. (10.3) wTiC was found to be constant within experimental error and no ternary term is needed. D. Calculation of the Molybdenum-Carbon and Chromium-Carbon Interaction Energies The values of wMoC and wCrC were calculated at x0 = 0 from the results in Table 9.1 for alloys 7262 and 726“, 7263 and 7265, and the following equations which are de- rived from Eq. (2.30). E -FCC-gr 6 = G C C + xN1“’N1C + + x XT1"’T10 MowMoC 2 _ 2 2 2 ’ xNixTiniTi ‘ leleleNl ‘ xNixMowNiMo ‘ xMoxNinoNi 2 2 ' xTixMowTiMo ' XMoxTiwmoTi, (10.8) —E _ —FCC-gr , GC ‘ G0 + XN1“’N10 + xTiniC + xCerrC 2 2 2 2 ’ xNixTiniTi ' xTixNiniNi ' xNixCrwNiCr ' xCrle‘chm 2 2 . ' xTixCeriCr ’ xCrxTiwCrTi’ (10'9) where ternary interaction energies are excluded. The previously calculated 013 were employed and the values of leCr’ wCrNi’ wNiMo’ I"Mom, wCrTi’ wTiCr’ wMoTi’wTiMo and 197 EEOC-gr were taken from Kaufman and Nesor (1973, 1975) (see Tables 10.2 and 10.5). The values of the interaction energies calculated assuming only binary terms were important are listed in Table 10.6. It is clear from the results in Tably 10.6 that in the cases of wMoC and wCrC that the calculated values are all composition dependent. Further, all of the binary interaction energies become more negative as the mole fraction of the total solute is increased. This means that the activity coefficient is smaller in the more concentrated solutions than would be expected from ex- trapolation of the dilute results. The trend is to lower than expected activity coefficients continued to an even larger extent in alloys 7267 and 7268 as discussed in Chapter IX. It thus appears that, as in the Fe-Ni-C system, the binary interaction energies are not sufficient to describe the systems in question. At xC = 0 the appropriate ternary terms from Eq. (2.30) are 6% (ternary) = Z 2 v i=1 J=1 n-1 n-2 n-1 2x x x +2 2 2- i=1 J=l k=1 (x1+xJ+xk k#1#J J .H.m cam N.m moanma CH panama n .ucoopoa Eoum ca co>fiw ohm mCOHpamanooo m.mm m.mm o.sm o.om oma sma Ha o.m + H2 woos m.mm o.sm s.Hm :.mm to =.m + o: H.: + no m.~ + H2 moms m.sm m.o= o.mo o.mo oz m.m + no 3.: + as m.m + a2 some 62 H.m + no m.m + as s.m + a2 some m.mm s.mm m.om m.mm oma omH no e.e + no m.m + H2 moms m.oe o.ms w.me m.me ooa men oz «.3 + as e.m + H2 some m.mm m.mm m.mm m.mm smH mHH no o.m + as :.m + H2 moms H.m= w.s= m.ms o.ms med sea oz m.m + as m.m + a2 moms H.Hs m.o= o.ao m.mm wma omH as s.m + H2 Home ooowm oHso noose oaoo noose oHoo snoaae mama OOHH com Aoov modumooQEoB Aoev ocoaofiooooo soa>fioo< .mHHmpsoeHnooxm oocfienopom osHm> one one coapmsum cmEmsmxlbofinox on» moans oopma Izoamo conpwo mo pCTAOHmmooo mpfi>fipo¢ on» no osam> on» no somfihwosoo .m.oa manms mumpoEmsma one .2.cH can m.cH moflnma :2 canon on one ncoauwasoano on» ad con: .opsmmoho oponomoepw H as oomph mononouoh pcocooEoo moon on» mean: oouwazoawo moaufi>auoOHH< oz< nc< 26< 02< hc< 29< 02< po< HB< mapa>duo< Acoc mucumpooEoB mama ccHH ccm .Aom.m .omc oofiossom oesoosxiooanox one means 6666226260 noooeoam weanoaae one no noaoa>2oo< .m.oa oases 206 10.9 have been used to calculate the solubility limit of carbon in the various alloys in equilibrium with pure titan- ium carbide according to _ -1 Ac ’ (ATin,TiC) where Kf TiC is the equilibrium constant for T1 (solv) + , i C (soln) = TiC (solid). Values of K are listed in f,TiC Table 8.“. TableilO.10comtains values of the activity of carbon at the titanium carbide solubility limit obtained from the results presented in Chapters VIII and IX and the values calculated with the Kohler-Kaufman equation. In this work the highest carbon activities investigated were 0.76 at 1215°C, 0.72 at 1100°C, and 0.59 at 900°C. Titanium carbide did not form, at any activity, in alloys 7261, 7265, and 7068. The lack of a two phase region, in these alloys, at the experimental activities is in agreement with the cal- culated solubility limit, in Table 10.10. In alloy 7263, which contains no molybdenum, the precipitate can be assum- ed to have an activity of one, based on arguments presented in Chapter VII. Experimentally, it is found that precipi- tation of titanium carbide does not commence, in alloy 7263, until an activity 50% higher at 900 and 25% at 1100 and 1215°C than the calculated value. This discrepancy could be due to experimental error. The data obtained from 2C)? Table 10.10. Comparison of Calculateda and Experimental Value of the Carbon Activity Where Precipitation of Titanium Carbide Should Start. Temperature/°C 900 1100 1215 AC Ac AC AC AC AC Alloy (Calc) (Exp) (Calc) (Exp) (Cale) (Exp) 7261 N1 + 2.“ T1 1.3 1.7 1.7 7262 N1 + 2.h Ti 0.33 0.17 0.56 0.18 0.63 0.18 +8.2 Mo 7263 N1 + 2.0 Tib 0.25 0.50 0.u3 0.50 0.u8 0.63 + 8.0 Cr ‘ 726” N1 + 2.“ Ti 0.66 0.36 0.99 0.38 1.0“ 0.32 + “.2 Mo 7265 N1 + 2.5 T1 O.L8 0.75 0.81 + u.6 Cr 7266 N1 + 2.“ Ti 0.06“ 0.1“ 0.0“6 0.19 0.067 + 8.3 Cr + 8.1 No 7267 N1 + 2.5 Ti 0.1a 0.27 0.19 0.32 0.095 + u.u Cr + 8.2 No 7268 N1 + 2.5 Ti 0.11 0.23 0.11 0.28 0.13 + 8.“ Cr + “.1 Mo 7068 N1 + 3.6 T1 0.61 0.36 0.91 aActivities calculated using results in Table 10.9 bExperimental values are approximate. They were obtained by interpolating between the solid solution data and one point in the two phase region. 208 Stover and Wulff (1959), although the best available, could be in error by 25% in the solubility product for titanium carbide. They relied on Curie point measurements, whose precision was not stated, to determine the phase boundary. Another possibility is that the model was not adequate. The assumption that titanium and nickel form a_temperature dependent regular solution in the nickel—rich corner of the phase diagram may be incorrect. Unfortunately, the true nature of Egi as a function of xT1 in nickel will have to await further data. Stover and Wulff's (1959) data do not cover a broad enough range of composition to yield more than one point on the dfii curve. TiC formed at all three temperatures in alloys 7262, 726“, 7266, 7267 and 7268. The solubility in these alloys determined experimentally is 1/3 to 1/2 the calculated solu- bility (Table 10.10). If the arguments in the preceding paragraph are correct the agreement between predicted and experimental solubilities are even worse. If one assumes that the molybdenum carbide forms an ideal solid solution with titanium carbide, the activity of the titanium, based on the compositions discussed in Chapter IX, would be 0.7 in alloy 7264 and 7268 and 0.58 in alloys 7262, 7266 and 7267. While lowering the activity of the carbide is a move in the right direction, the change is not sufficient to bring the calculated and observed values together. The most plaus- ible explanation for the remaining discrepancy is that, 209 rather than forming an ideal solution, the carbides mix with a negative heat of mixing. If a value of approximately -6.7 i 2 kJ°mol-1 is assumed for the heat of mixing and if the entropy of mixing is assumed to be ideal, the calculated and experimental values of the solubility agree to :15 per- cent. A slightly more negative value for the heats of mixing is needed if the calculated values of AC are shown to be too low. Clearly, more precise thermodynamic data are required for the nickel-titanium system in order to resolve the discrepancies. CHAPTER XI THERMOMIGRATION A. Introduction Until recently, thermomigration, the mass flux induced by a temperature gradient, was studied exclusively in liquids and gases. Experimental difficulties associated with establishing and maintaining a large, well-defined tempera- ture gradient in a solid dissuaded researchers from investi- gating thermomigration in solids. Modern work in the field started with Shewmon (1958) and Darken and Oriani (195“) who investigated several metal-metal and metal-metalloid systems. Oriani (1969) reviewed the 1960's experiments on metal-metalloid binary systems, which yielded little quan- titative data. Poor temperature control and poor chemical analyses plagued most investigators. Thermomigration in solids is an important phenomenon in, for example, nuclear reactors and in welding. In nuclear reactors, large temperature gradients are the norm rather than the exception. Thermomigration of hydrogen in the Zircalloy fuel cladding and in the oxide fuel are of great technical importance. In welding the tremendous tempera- ture gradients at the liquid-solid interface cause a mass flux which may be responsible for cracks that form in many 210 211 welds after cooling. Thermomigration experiments have as their immediate goal the measurement of the "thermal diffusion factor", 01. For a binary system with a linear temperature gradient in the Z direction, 01 can be determined from [Horne and Anderson (1970)] WI = alw1w2[l + H—-exp(— —t/0)]sin(T n) (11.1) wl = weight fraction of component i w; = initial weight fraction of component i Z = coordinate in the direction of the temperature gradient. At the center of the specimen Z = 0. t = time d = the diffusion pathlength e = d2/n2D relaxation time D = binary diffusion coefficient Equation (11.1) indicates that the composition of the specimen as a function of position will continue to change until t 3 us, after which time a steady state will persist as long as the temperature gradient is maintained. Measure- ments made after t = H0 will not provide any information on D but do provide data for calculation of 01. To date the few thermomigration experiments in solids have all been done at the steady state (t 5 he). In this work the 212 measurements were to be time dependent so that al and D could be determined in the same experiment. B. Experiments Specimens were annealed in a temperature gradient of ap- proximately 1000°C/cmixia.Gleeble. A Gleeble is an instru- ment designed to simulate the large temperature fluctuations produced in metal alloys during welding. A cylindrical sample is clamped at both ends in water cooled copper Jaws, and a large alternating current is then passed through the sample. The sample is brought from 20 to 1300° C in less than 10 sec. The temperature of the sample is controlled via a feedback loop containing a thermocouple attached to the center of the sample. Solution of the heat conduction equation for this experimental arrangement as well as actual experimental measurements show that the temperature distribution in the sample is parabolic with a maximum in the center. For sample B—6-B the temperature was found to obey T: = .2371d2 + 82.06d + 1350. with the root mean square residual c = 10°C. The tempera- ture of the sample was measured at three sites on the specimen with platinumeplatinum 10% rhodium thermocouples 213 news pswas mes .mcou pos ea soap iwuahsnhmoov mpoz .coccma zaopmsaxOAQQm use no: on» ma Oposa mo 06am im .oanomac on» :a mason 03» cwamchm xmm pm m in oaQEmm .a aa opsmam 21“ and recorded as voltage on a three pen pentiometric strip- chart recorder. The end temperatures were also known. The atmosphere around the samples was supposed to be con- trolled by flowing pure argon at approximately 100 liters per hour through a pyrex cover box surrounding the sample. C. Results Several specimens were annealed in the Gleeble for times varying from five minutes to two hours. The results were of two kinds: either a gradient of carbon concentration was not observed or the sample was partially decarburized. Figure]J:J.shows half of a sample annealed two hours in the Gleeble and then annealed 100 hours at 760°C to pre- cipitate the carbon from solution. The carbon distribution in the sample approaches the shape of an hour glass. This distribution would be expected in a sample with a sink at the surface and a maximum in temperature at the center. From these results it is apparent that better control over the atmosphere surrounding the sample is necessary if quantitative results are to be obtained. Cost, time con- sideration, and the requirements of other users mitigated against modification of the Gleeble for further study of thermomigration. . There is still a need for thermomigration experiments in interstitial metal alloys, and a suitably modified 215 Gleeble would offer many advantages, such as rapid heat-up and cool-down. The modification most needed is a high quality vacuum system in order to control the chemical environment surrounding the specimen. CHAPTER XII SUGGESTIONS FOR FURTHER WORK A. Analytical Chemistry While a great deal of effort has been expended in im- proving techniques for analysis, further improvements are still desirable. The carbon analyses are in need of ac- curate standards; as discussed in Chapter III, the standards currently available have an accuracy of about 15%. The carbon analyses could also be improved if a more selective detector were used. Our apparatus used a conductometric detector. Newer instruments use infra-red detectors, which are not as sensitive to impurities such as SO2 and do not require CO2 traps and chromatographic columns. In the area of metal analysis more study is needed on "matrix" effects in the acidic solutions. These effects require the use of standards of similar composition to the samples. In some cases this is not convenient or possible. For analysis of small quantities of solid material the de- velopment of x-ray fluorescence capability would be desir- able. The electron microprobe technique, while useful, is limited in that only relative concentrations are readily obtainable. 216 217 B. Experiments Six different series of alloys need to be studied in order to understand better the ternary interactions that this research has revealed. The six systems are Ni—Mo—C, Fe-Mo-C, Ni-Cr-C, Ni-Mo-Cr-C and Fe-Mo-Cr-C. Experiments should be carried out with as high a concentration of Mo and/or Cr as possible without leaving the faceacentered cubic solid solution phase field. The goal of these experi- ments would be to determine quantitatively the values of the ternary interaction energies. The question of whether there is any solvent dependence in the binary interaction energy could also be reSolved by these experiments. If the binary interaction energies determined in nickel and iron solutionsckanot agree once ternary terms are taken into account, still higher order terms will have to be introduced into Kohler-Kaufman formalism. In solutions with low carbon solubility the car-' burization technique needs to be refined to facilitate experiments at carbon activities of less than 0.05. This would involve using gas mixtures of lower CHu/Hg ratios and possibly lowering P02 in the furnace. The result would be a better understanding of the titanium-molybdenum- carbon precipitation process and the molybdenum-chromium- carbon solid solution interaction. More controlled experiments are necessary on the 218 precipitation of carbon upon quenching. Resistance heating and a helium quench offer the most convenient methods of controlling the quench rate. Annealing samples at tempera- tures of around 500°C for short periods of times and observ- ing changes in the weight percent of the precipitate and in the x-ray diffraction patterns would provide insight into the precipitation process. It is also hoped that short anneals at low temperatures would allow the precipitates to grow large enough to be viewed in the electron micro- scope. APPENDIX A APPENDIX A The compositionscfi‘the uncarburized alloys are given in Table A-1. Tables A-2 through A-33 contain all of the gas phase carburization data generated in this investigation. The data in each table constitute one data set. That is, all of the specimens in the set were carburized at the same time in the same furnace run. Thus, the temperature and the equilibrating gas are identical for all the specimens described in a given table. For these two reasons all are listed together. Unfortunately, the analytical standards used for carbon analysis of the specimens, even in a specific table, are not all the same. This arose because the supply of National Bureau of Standards Standard Reference Material (NBS SRM) 19E was exhausted. Thus, when rechecking specimens in some tables, a different NBS SRM was used. (In some tables, of course, only one NBS SRM was used.) As discussed in Chapter III, it is important when using the carbon data to relate all of the concentrations to the same NBS SRM. In all of the calculations in this work the carbon concentrations are rela- tive to NBS SRM 19E. Extensive comparison of SRM 19E and 121B (the only other standard used in the carbon analyses) showed that a concentration relative to 121B must be multi- plied by 0.966 to obtain the concentration relative to 19E. Analytical carbon data were rechecked frequently, as is 219 220 partially apparent from examination of the variation of NBS SRM's in the tables. Shortly after the gas phase carburiza- tion studies began it was apparent that problems existed in our ability to analyze for carbon. Comparison of weight change and the carbon analysis did not always agree. In- consistencies between data sets and the size of the aliquot used in the analysis affected the results. Once it was realized that analytical difficulties existed, the stringent controls on the combustion procedure detailed in Chapter IV were developed. Unfortunately, before all of the analytical problems were solved, the supplies of four sets of specimens, A—7603-118, A-7603-l2l, A—7783-20, and A-7783-21, had been exhausted. When these specimens were analyzed the instru- ment was giving consistently low values for the carbon concentration when small aliquots were used. In the four sets of specimens mentioned above all of the one phase specimens, except the iron standards, contained less than 0.05 wt. % carbon. Analysis of data from these specimens showed that they had uniformly low activity coefficients relative to samples analyzed after the instrument problems had been corrected. In the final analysis of the data, therefore, the activity coefficient of the nickel alloys in the aforementioned data sets was obtained from the ac— tivity coefficient of carbon in nickel determined in the data sets listed in Table 5.1. The carbon analyses of specimens Ni-A7783—16 and 7068-A7603-106 were disregarded 221 in the analysis of the results. In both specimens the cal- culated activity coefficients were more than 3 standard deviations from the mean value and were not consistent with the other data sets with respect to equilibrium concentra- tions of carbon. That is, in data set A-7783—l6 the nickel specimen analyzed to be lower in carbon than N1 + A at. % Mo (726A) and in A-7603-106 alloy 7068 analyzed to be lower in carbon than nickel. These are contrary to the results of all the other data sets. Data set A-7783-36 has not been considered in the analysis of the data. Repeated analyses of the specimen from this set gave non-repro- ducible results even when the carbon analyzer appeared to be functioning properly. The abbreviation T.P. in the tables indicates that the specimen was assumed to be two phase, although the material was not extracted. The specimens were Judged two phase on the basis of their activity coefficients. A decrease in the carbon activity coefficient at high carbon activity indicates that precipitation has occurred. 222 a Table A.l. Composition of Alloys Used for Calculations. Alloy Element/wt % MElt 2 Nb. T1 Cr Mb C N1 7261 2.0 0.015 98.0 7262 1.95 12.81 0.01“ 85.3 7263 2.02 7.78 0.01“' 90.2 726“ 1.95 6.68 0.015 91.4 7265 2.06 “.09 0.016 93.8 7266 1.91 7.08 12.76 0.021 78.2 7267 1.95 3.77 12.93 0.016 81.“ 7268 2.00 7.33 6.66 0.015 8U.0 7071 2.8 8.08 0.135 89.0 7095 3.06 13.9 0.380 82.7 A 2.0 13.0 0.09“ 8U.9 B 1.73 0.086 98.2 C 2.0 7.UO 0.109 90.5 14149 1.95 7.36 11.14 0.035 79.0 8These values were picked from those in Table 5.1. 223 Table A.2. Data From Carburization Experiment A-7603-97. Date: 4/28/76; Temperature: 1215°C; Duration: 40 hours H20(g) Concentration: 1.5 ppm; Quench: Water. Final Microprobe [C] Cal. Intensity Weight (wt %) Std. Ratio Lattice Change by For Precip . (Mo ) Parameter Alloy ( %) Analysis Carbon (wt . % ) IT aO/nm 7261 0.124 0.135 19E 0.24 7262 0.325 0.321 121B 1.57 1.50:0.03 0.43158 7263 0.132 0.160 1213 0.25 7266 0.832 0.852 19E 6.05 3.00 0.08 Ni 0.128 0.131 19E 0.22 aoa =0.0001 nm. 0 Table A.3. Date: 1215°C; Duration: Data From Carburization Experiment A27603-105 5/ 4/ 76 3 Temperature : 36 hours; H20(g) Concentration: 1 ppm; Quench: Cold Zone. Final Microprobe [C] Cal. Intensity weight (wt %) Std. Ratio Lattice Change by For Precip. (Mb) Parameter Alloy (% ) Analysis Carbon (wt . %) TI ao/nm 7264 0.125 0.138 19E 0.147 7265 0.147 0.162 19E 7267 0.687 0.708 1213 TP 7268' 0.487 0.523 121B TP Ni-270 0.146 0.150 121B 22“ Table A.4. Data From Carburization Experiment A27603-106. Date: 5/6/76; Temperature: 1215°C; Duration: 36 hours; H20(g) Concentration: 1 ppm; Quench: Cold Zone. Final ‘Microprobe [C] Cal. Intensity Weight (wt Z) Std. Ratio Lattice Change by For Precip . (1&1) Parameter Alloy (1) Analysis Carbon (wt. %) Ti aO/hm 7068 0.0663 0.140 19E 7071 0.0802 0.217 19E 7095 0.477 0.870 19E 6.19 1.75i0.02 'Ni-270 0.140 0.150 19E 0.24 Ni-270 0.110 0.147 19E Fe'E' 0.941 0.981 19E 225 Table A.5. Data From Carburization Experiment A-7603-118. Date: 5/17/76; Temperature: 1215°C; Duration: 22 hours; H20(g) Concentration: 1 ppm: Quench: Cold Zone. Final Nficroprobe [0] Cal. Intensity Weight (wt %) Std. Ratio Lattice Change by For Precip . (Mg) Parameter Alloy (%) Analysis Carbon (wt. %) Ti ao/nm 7261 0.020 0.0337 19E 7262 0.053 0.0274 19E 7263 0.038 0.0410 19E 7264 0.086 0.0285 19E 7265 0.040 0.0359 19E 7266 0.059 0.0557 19E TP 7267 0.047 0.0358 19E 7268 0.103 0.0414 19E 7068 0.188 0.0355 19E Ni-270 0.023 0.0330 19E Fe'E'a 0.261 121B ‘ 0.256 19E ainitial wt not recorded. 226 Table A.6. Data From Carburization Experiment A-7603-121. Date: 5/18/76; Temperature: 1215°C: Duration: 46 hours; H20(g) Concentration: 1 ppm; Quench: Cold Zone. Final Nflcroprobe [0] Cal. Intensity Weight (wt %) Std. Ratio Lattice Change by For* Precip. (M60 Parameter Alloy (%) Analysis Carbon (wt. %) Ti ao/nm 7261 0.035 0.0442 19E 7262 0.020 0.0337 19E 7263 0.019 0.0482 19E 7264 0.013 0.0364 19E 7265 0.018 0.0447 19E -7266 0.124 0.142 19E 0.639 0.43113 7267 0.044 0.0461 19E 7268 0.019 0.0535 19E 7068 -0.044 0.0444 19E 7095 -0.301 0.0795 19E 7071 -0.067 0.0568 19E Ni-270 0.041 0.0433 19E Fe'E' 0.291 0.325 19E 0.330 1218 ac =0.0001 nm. Table A.7. Date: 227 Data From Carburization Experiment A27603-123. 5/20/76; Temperature: 215°C; Duration: 64 hours; H20(g) Concentration: 1 ppm; Quench: Cold Zone Final Microprobe [C] Cal. Intensity Weight (wt z) Std. Ratio Lattice Change by For Precip. (M90 Parameter Alloy (%) Analysis Carbon (wt. %) Ti aO/nm 7261 0.367 0.398 1218 7263 1.91 0.535 121B 0.963 7264 0.683 0.675 121B 2.55 1.15:0.03 .4326a 7265 0.441 0.447 121B 7268 1.058 1.085 19E TP 7068 0.377 0.449 121B 7071 0.657 0.832 121B Ni-270 0.355 0.37 19E a Ca =0.0001 nm. 0 Table A.8. Data From Carburization Experiment A-7783-4. Date: 6/16/76; Temperature: 1100°C; Duration: 48 hours; H20(g) Concentration: 1 ppm; Quench: Gold Zone. Final Nflcroprobe [C] Cal. Intensity Weight (wt %) Std. Ratio Lattice Change by For Precip. (E20 Parameter Alloy (%) Analysis Carbon (wt %) Ti aO/nm 7261 0.024 0.0407 19E 7262 0.027 0.0323 121B 7263 0.043 0.0450 121B 7266 0.185 0.190 1213 1.36 1.2410.02 .4311a Ni-270 0.029 0.0376 1213 F'E' 0.324 0.355 121B aoa =0.0001 nm. 0 Table A.9. 228 Data From Carburization Experiment A-7783-14 Date: 7/1/76; Temperature: 1100°C: Duration: 48 hours; H20(g) Concentration: 2.5 ppm; Quench: Cold Zone. Final Microprobe [C] Cal. Intensity Weight (wt 7:) Std. Ratio Lattice Change For Precip. (M90 Parameter Alloy (%) Analysis Carbon (wt. %) Ti aO/nm 7264 0.017 0.0363 19E 7265 0.019 0.0452 19E 7267 0.071 0.0852 19E 0.385 1.171001 7268 0.048 0.0577 1218 TP Ni-270 0.032 0.0381 19E Fe'E' 0.350 0.401 1218 Table A.10. Data From Carburization Experiment A-7783-l5. Date: 7/3/76; Temperature: 1100°C; Duration: 72 hours; H20(g) Concentration: 1 ppm; Quench: Cold Zone. Final Nficroprobe [C] Cal. Intensity Weight (wt %) Std. Ratio Lattice Change by For Precip. (M20 Parameter Alloy (%) Analysis Carbon (wt. %) Ti ao/nm 7262 0.330 0.346 1218 2.14 1.29:0.02 7266 0.758 0.835 121B 6.53 2.44:0.06 .7267 0.561 0.610 1218 3.99 1.44:0.01 7268 0.424 0.467 1218 2.39 0.82:0.02 Ni-270 0.090 0.0951 121B Fe'E' 0.733 0.796 1218 Table A.11. Date: Data From Carburization Experiment A-7783-l6 7/6/76; Temperature: 1100°C; Duration: 48 hours; H20(g) Concentration: 1 ppm; Quench: Cold Zone. Final ' Microprobe [0] Cal. Intensity - Weight (wt %) Std. Ratio Lattice Change by For Precip. (M20 Parameter Alloy (%) Analysis Carbon (wt. %) Ti ao/nm 7261 0.081 0.0959 1218 7263 0.118 0.116 1218 7264 0.065 0.082 1218 TP 7265 0.090 0.106 1218 Ni-270 0.077 0.0793 1218 Fe'E 0.784 0.811 1218 Table A.12. Date From Carburization Experiment A-7783-17 Date: 7/8/76; Temperature: 1100°C; Duration: 46 hours; H20(g) Concentration: 1 ppm; Quench: Cold Zone. Final 'Microprobe [C] Cal. Intensity Weight (wt 3) Std. Ratio Lattice Change by For Precip. (M29 Parameter Alloy' (%) Analysis Carbon (wt. %) Ti ao/nm 7261 0.210 0.244 1218 7263 0.345 0.383 1213 0.99 0.43248 7264 0.456 0.506 121B 2.14 0.91:0.03 0.43218 7265 0.228 0.252 1218 Ni -270 O. 208 0.225 1218 Fe'E' 1.424 1.49 19E BO’a =0.0001 nm. 0 Table A. 13 0 Date : 1100°C; Duration: Data From Carburization Emeriment A-7783-l8. 7/10/76; Temperature: 48 hours; H20(g) Concentration: 1 ppm; Quench: Cold Zone. Final Microprobe [0] Cal. Intensity Weight (wt %) Std. Ratio Lattice by For Prec ip . (119) Parameter Alloy (%) Analysis Carbon (wt. %) Ti aO/nm 7261 0 . 1614 0 . 202 1218 7263 0 . 227 O . 235 19E 72614 0 . 1814 0 . 325 19E TP 7265 0 . 314 0 . 196 19E 7068 0.1147 0.222 1218 Ni-270 0 . 161 0 . 179 198 Fe"E' 1.20 1.28 1213 Table A.1Ll. Data from Carburization Experiment A-7783-19. Date: 7/13/76; Temperature: 1100°C; Duration: 48 hours; H20(g) Concentration: 2 ppm; Quench: Cold Zone. Final Microprobe [C] Cal. Intensity weight (wt %) Std. Ratio Lattice Change by For Precip . (gig) Parameter Alloy (%) Analysis Carbon (wt. Z) Ti ao/nm 7262 0.160 0.174 1213 0.945 1.27:0.02 0.4318a 7266 0.556 0.6214 121B 3.84 1.48:0.04 7267 0.431 0.460 193 2.95 1.33:0.02 0.43148 7268 0.283 0.310 1213 1.52 0.85:0.03 0.43198 7068 0.00 0.0873 1218 Ni-270 0 . 067 0 . 0726 19E Fe'E' 0.577 0.616 1218 a 0 Ga =0.0001 nm. 231 Table A.15. Data From Carburization Experiment A27783-20. Date: 7/16/76; Temperature: 1100°C; Duration: 60 hours; H20(g) Concentration: 2 ppm; Quench: Cold Zone. Final Microprobe [C] Cal. Intensity Weight (wt 5) Std. Ratio Lattice Change by For Precip. (Mb) Parameter Alloy (Z) Analysis Carbon (wt. %) TI' aO/nm 7262 0.014 0.0182 1218 7266 0.064 0.0633 19E 7267 0.033 0.0228 19E 7268 0.033 0.0285 19E 7068 -0.063 0.0253 19E Ni-270 0.0128 0.0222 19E Fe'E' 0.179 0. 205 1218 Table A.16. Data From Carburization Experiment A-7783-21 Date: 7/16/76; Temperature: 1100°C; Duration: 60 hours; H20(g) Concentration: 2.5 ppm; Quench: Cold Zone. Final Microprobe [C] Cal. Intensity Weight (wt %) Std. Ratio Lattice Change by For Precip . ( Mg) Parameter Alloy (%) Analysis Carbon (wt. 1) Ti ao/nm 7262 0.00 0.0106 198 7266 0.02 0.0192 19E 7267 0.01 0.0133 198 7268 0.00 0.0164 19E Ni-270 0.00 0.0140 19E , Fe'E' 0.084 0.111 19E 232 Table A.17. Data From Carburization Experiment A-7783-32. Date: 7/22/76; Temperature: 1100°C; Duration: 85 hours H20( g) Concentration: 1 ppm QUench: Cold Zone. r Final Microprobe [C] Cal . Intensity Weight (wt 1) Std. Ratio Lattice Change by For Precip. (Mo ) Parameter Alloy (%) Analysis Carbon (wt . 1) TI ao/nm Ni-270 ‘3 -0.177 0.169 19E Ni-270 0 . 154 0 . 174 1218 Fe'E'a -0.271 1.206 19E Fe'E l. 157 1. 191 19E 7068 0 . 116 0 . 204 1218 aEquilibrium approached by decarburization. Table A.18. Data from Carburization Experiment A-7783-33. Date: 7/24/76; Temperature: 1100°C; Duration: 60 hours; H20( g) Concentration: 1 ppm; Quench: Gold Zone. Final Mi croprobe [C] Cal . Intensity Weight (wt 1 ) Std. Ratio Lattice Change by For Precip . (Mo ) Parameter Alloy (%) Analysis Carbon (wt. 1) “TI ao/nm Fe'E' 1.076 1.114 19E Table A.19. 233 Data From Carburization Experiment A27783-35. Date: 7/29/76; Temperature: 1100°C; Duration: 90 hours; H20(g) Concentration: 1 ppm; Quench: Gold Zone. Final Microprobe [C] Cal. Intensity Weight (wt 7) Std. Ratio Lattice Change by For Precip. (Mb) Parameter Alloy (1) Analysis Carbon (wt. 1) TT' aofimm 7068 0.102 0.174 1218 7264 0.170 0.177 19E 0.475 0.84:0.04 0.4324a Ni-270 0.125 0.136 19E Fe'E' 1.008 1.032 19E 8 _. Ga -0.0001 nm. 0 Table A.20. Data From Carburization Experiment A-7783-36. Date: 8/2/76; Temperature: 1215°C; Duration: 96 hours; H20(g) Concentration: 1 ppm; QUench: Gold Zone. Final Microprobe [0] Cal. Intensity Weight (wt 7) Std. Ratio Lattice ' Change by For Precip. (Mo) Parameter Alloy ,(3) Analysis Carbon (wt. 1) TI' aO/nm 7261 0.152 0.183 1218 7263 0.191 0.250 1218 7264 0.146 0.181 1218 TP 7265 0.172 0.212 1218 7068 0.116 0.217 121B Ni-270 0.150 0.174 1218 Fe'E' 1.01 1.18 1218 234 Table A.21. Data From Carburization Experiment A—7783-37. Date: 9/8/76; Temperature: 1215°C; Duration: 108 hours; H20(g) Concentration: 1 ppm; Quench: Cold Zone. Final Microprobe [C] Cal . Intensity Weight (wt 1) Std. Ratio Lattice Change by For Precip . (Mo ) Parameter Alloy (%) Analysis Carbon (wt . 1) TI ao/nm 7262 0.415 0.462 1213 2.17 1.55:0.04 0.43133 7266 0.865 0.985 1213 7.20 3.19:0.04 0.42998 7267 0.670 0.677 19E 4.42 1.90:0.05 0.43038 7268 0.448 0.512 1213 2.11 0.92:0.03 0.4316a Ni-270 0. 107 0 . 157 1218 Fe'E' 0.961 1.03 1218 a da =0.0001 nm. 0 Table A. 22 . Data From Carburization Experiment A-778 3— 38 . Date: 9/14/76; Temperature: 1215°C; Duration: 60 hours; H20( g) Concentration: 1 ppm; Quench: Cold Zone . Final Microprobe [C1 Cal . Intensity Weight (wt 1) Std. Ratio lattice Change by For Precip . (Mo ) Parameter Alloy ( % ) Analysis Carbon (wt . 1 ) TT ao/nm 7262 0 . 088 0. 0971 1218 TP 7266 0.526 0.626 1213 3.62 1.91:0.06 0.43008 7267 0.323 0.384 1213 1.85 1.50:0.02 0.43108 7268 0.146 0.202 1213 0.394 0.75:0.02 0.431b Nil-270 0. 084 0. 0932 1218 Fe 'E' 0. 624 0.750 1218 aoa =0.0003 mn- o boa =0.0001 nm. 0 235 Table A.23. Data From Carburization Experiment A-7783-44. Date: 10/9/76; Temperature: 900°C; Duration: 108 hours; H20( g) Concentration: 1 ppm; Quench: Cold Zone. Final. Microprobe [C] Cal. Intensity Weight (wt 1) Std. Ratio Lattice Change by For Precip. (1142 Parameter Alloy (Z ) Analysis Carbon (wt . X ) Ti ao/nm 7261 0. 075 0 . 100 1218 7263 0.195 0.218 1213 0.885 0.43208 7264 0.206 0.235 1218 1.342 0.90 0.02 0.43268‘ 7265 0. 086 0 . 0987 1218 7068 0 . 031 0 . 107 1218 Ni-270 0.074 0.0953 121B Fe'E' 0.781 0.832 1218 '87 0a =0 . 0001 nm. 0 Table A.24. Data From Carburization Experiment A-7783-45 Date: 10/16/76; Temperature: 900°C; Duration: 132 hours; H20(g) Concentration: 0.5 ppm; Quench: Cold Zone Final Microprobe [C] Cal . Intensity weight (wt 7) Std. Ratio Lattice Change by For Precip . (Mo ) Parameter Alloy (% ) Analysis Carbon (wt . Z ) TI ao/nm 7261 0.035 0.0603 1213 7263 0.038 0.0634 1213 7264 0.011 0.0580 1213 7265 0.037 0.0624 1213 7068a -0.031 0.0620 1213 Ni-270 0.049 0.0544 1213 Ni-27(? -0.058 0. 0538 1213 Fe'E' 0.501 0.544 1213 aEquilibrium approached by decarburization. 236 Table A.25. Data From Carburization Experiment A-7783-47. Date: 10/23/76; Temperature: 900°C; Duration: 120 hours; H20( g) Concentration: 0.5 ppm; Quench: Gold Zone. Final Microprobe [C] Cal . Intensity Weight (wt 1) Std. Ratio Lattice Change by For Precip . (Mo ) Parameter Alloy ( Z ) Analysis Carbon (wt . Z) TI ao/nm 7261 0.024 0.0457 1218 7263 0.032 0.0445 1218 7264 0.026 0.0360 1218 7265 0.019 0.0436 1218 7068 -0.036 0.0505 1218 Ni-270 0.033 0.0425 1218 Fe'E' 0.407 0.444 1218 Table A.26. Data From Carburization hperiment A-7783-48. Date: 10/28/76; Temperature: 900°C; Duration: 120 hours; H20( g) Concentration: 0.5 ppm; Quench: Cold Zone. Final Microprobe [C] Cal . Intensity Weight (wt 1) Std. Ratio Lattice Change by For Precip. (Mo ) Parameter Alloy ( Z) Analysis Carbon (wt . Z) T1 ao/nm 7262 0 . 083 0 . 10 2 1213 7266 0.083 0.081 1213 1.22 1.64:0.02 7267 0.202 0.243 1213 0.582 1.2510.02 0.43183 7268 0.176 0.191 1213 1.56 0.94:0.02 0.4320a 31-270 0.027 0.0423 1213 Fe'E' 0.380 0.411 1213 aoa =0.0001 nm. 0 Table A.27. 237 Data From Carburization Experiment A-7783-49. Date: 11/3/76; Temperature: 900°C; Duration: 120 hours; H20( g) Concentration: 0.5 ppm; Quench: Cold Zone. Final Microprobe [0] Cal . Intensity Weight (wt Z) Std. Ratio Lattice Change by For Precip. (Mg) Parameter Alloy (Z) Analysis Carbon (wt. Z) Ti ao/nm 7262 0 . 007 0 . 0125 1218 72 66 0 . 007 0 . 0218 1218 7267 0.040 0.0458 1213 0.411 1.21:0.003 0.43138 7268 0 . 021 0 . 0349 1218 0. 042 Ni-270 0.0141 0.0166 1218 Fe'E' 0.168 0.191 1218 a .. 0a -000ml nm. 0 Table A.28. Data From Carburization Experiment A-7783-57. Date: 11/9/76; Temperature: 900°C; Duration: 144 hours; H20( g) Concentration: 0.5 ppm; Quench: Gold Zone. Final Microprobe [C] Cal . Intensity Weight (wt 1) Std. Ratio Lattice Change by For Precip . (Mo ) Parameter Alloy (Z) Analysis Carbon (wt. Z) Ti- ao/nm 7262 0 . 0226 0 . 0326 1218 7266 0 . 0186 0 . 0362 1218 7267 0 . 0598 0 . 0465 1218 7268 0 . 0973 0 . 112 1218 Ni-270 0.0189 0.0299 1218 Fe'E' 0.287 0.312 1218 238 Table A.29. Data From.Carburization Experiment A27783-116. Date: 1/22/77; Temperature: 1215°C; Duration: 48 hours; H20(g) Concentration: 0.5 ppm; Quench: Cold Zone Final Microprdbe [0] Cal. Intensity Weight (wt 1) Std. Ratio Lattice Change by For Precip. (Q) Parameter Alloy ( Z) Analysis Carbon (wt . Z ) Ti aO/nm 7261 0.064 0.0747 121B 7262 0.079 0.0628 1213 7263 0.065 0.0862 1213 7264 0.062 0.0640 1218 7265 0.057 0.0810 121B 7266 0.413 0.475 121B 7267 0.198 0.221 1218 T? 7268 0.072 0.109 1218 TP 7068 0.002 0.0783 121B Ni-270 0.068 0.0704 1218 Fe‘E' 0.489 0.554 1218 Table A.30. 239 Data From Carburization Experiment A-7783-120. Date: 1/25/77; Temperature: 1215°C; Duration: 60 hours; H20( g) Concentration; Quench: Cold Zone. I Final Microprobe . [C] Cal . Intensity Weight (wt 1) Std. Ratio Lattice Change by For Precip. (Mo ) Parameter Alloy (Z) Analysis Carbon (wt. Z) TI ao/nm 7261 0.132 0.147 1218 7262 0.352 0.361 1218 TP 7263 0.142 0.173 1218 7264 0.127 0.124 1218 7265 0.146 0.153 1218 7266 0.781 0.867 1218 7267 0.598 0.652 1218 TP 7268 0 . 423 0. 4 51 1218 T? 7068 0 . 084 0 . 1 52 1218 Ni-270 0 . 150 0 . 144 1218 Fe'E' 0.911 1.01 1218 240 Table A.31. Data From Carburization Experiment A.--7783-123a Date: 2/8/77; Temperature: 1215°C; Duration: 72 hours; H20(g) Concentration: 0.5 ppm; Quench: Cold Zone. Final [C] Cal. Weight Initial (wt 1) Std. . Change [C] by For Precip. Alloy (Z) (wt Z) Analysis Carbon (wt. Z) 7261 +0.020 0.147 0.137 1218 7262 +0.001 0.361 0.346 1218 TP 7263 -0.027 0.173 0.163 1218 7264 +0.003 0.124 0.125 1218 7265 -0.085 0.153 0.143 1218 7266 -0.034 0.867 0.859 1218 TP 7267 -0.017 0.652_ '0.634 1213 TP 7268 -0.055 0.451 0.423 1218 -TP 7068 -0.005 0.152 0.143 1218 Ni-270 +0.008 0.144 0.131 1218 Fe'E' -0.106 1.01 0.959 1218 aEquilibrium approached by decarburization. Table A.32. Date: Data From Carburization Experiment A.-7783-125.a 2/11/77 ; Temperature: H20(g) Concentration: 0.5 ppm; Quench: Gold Zone. 241 1100°C; Duration: 96 hours; Final [C] Cal. Weight Initial (wt Z) Std. Change [C] by For Precip. Alloy (Z) (wt. 5:) Analysis Carbon (wt. 1) 7261 -0.071 0.147 0.0691 1213 7262 -0.126 0.361 0.218 1213 TP 7263 -0.074 0.173 0.0745 1213 7264 -0.069 0.124 0.0570 1213 7265 -0.058 0.153 0.0714 1213 7266 -0.096 0.867 0.763 1213 TP 7267 —0.110 0.652' 0.515 1213 TR 7268 -0.135 0.451 0.316 1213 TR Ni-270 -0.076 0.144 0.0643 1213 Fe'E' -0.352 1.01 0.565 1213 aEquilibriumapproached by decarburization. 242 Table A.33. Data From Carburization Experiment A.-7783-136.a Date: 2/17/77; Temperature: 990°C; Duration: 192 hours; H20(g) Concentration: 0.5 ppm; Quench: Gold Zone. Final [C] Cal. Weight Initial (wt 1) Std. Change [C] by For Precip. Alloy (Z) (wt. Z) Analysis Carbon (wt. Z) 7261 -0.016 0.075 0.049 1213 7262b -0.002 0.218 7263 -0.006 0.087 0.0847 1213 7264 -0.011 0.064 0.0537 1213 7265 -0.030 0.081 0.0472 1213 7266b +0.004 0.763 7267b - +0.009 0.515 7268b +0.003 0.316 7068b -0.015 0.078 0.0578 1213 Ni-270 -0.030 0.070 0.0447 1213 Fe'E' -0.113 0.553 0.462 1213 aEquilibrium approached by decarburization. quuilibrium.was not achieved in these alloys. BIBLIOGRAPHY BIBLIOGRAPHY M. Abdel-Gawad, The American Mineralogist, 51, 464 W. Andrews and H. Hughes, Am. Soc. Testing Mats., ASTM, STP 393, 3 (1966). Ansara, Metallurgical Chemistry Symposium, 0, Kubaschew- ski editor, National Physical Society, 403 (1971). . Ban-Ya, J. F. Elliott, and J. Chipman, Met. Trans., 1, 1313 (1970)- Ban-Ya, J. F. Elliott, and J. Chipman, TMS-AIME, 245, 1199 (1969). Beck and R. Bigot, Rev. Met. (Paris) 62, 681 (1965). Belzer, L. G. Savedoff, and H. L. Johnston, Ohio State U. Research Foundation TR316-6 (1953). Benz and J. F. Elliott, TMS-AIME, 215, 954 (1959). . J. Bradley, M. 0. Pattengill, and L. M. Ferris, Inor- ganic Chemistry, 4, 1080 (1965). D. N. Braski and J. M. Seitnaker, Production of Homogen- eous Titanium-Hastelloy N Alloys, 0RNL/TM-5697 (Feb. 1977). K. Bungardt, H. Preizendanz, and G. Lehnert, Arch Eisen- huett., 35, 999 (1964). R. A. Buckley and W. Hume-Rothery, J. Iron Steel Inst., J. J. MLOC-lc-i 291, 227 (1963). Chipman, Met. Trans., 3, 55 (1972). Chipman, TMS-AIME, 239, 1332 (1967). Chipman, TMS-AIME, 232, 2 (1967). Chipman, J. Iron and Steel Institute, London, 180, 97 (1955). Chipman and E. F. Brushy, TMS-AIME, 242, 35 (1968). 0. Clayton and W. F. Giaugue, J.A.C.S., 24, 2610 (1932). 243 244 L. S. Darken, JACS, 33, 1163, Appendix (1946); ” L. S. Darken, TMS-AIME, 139, 430 (1949). L. S. Darken, TMS-AIME, g39, 80 (1967). L. S. Darken and R. A. Oriana, Acta Met. 2, 841 (1954). M. J. Donachie, Jr. and 0. H. Kriege, Journal of Materials, 1, 269 (1972) C. K. H. DuBose and J. 0. Stiegler, Rev. Sci. Inst., _8, 694 (1967). W. W. Dunn, R. B. McLellan, and W. A. Oates, TMS-AIME, 242, 2129 (1968). H. Dunwald and C. Wagner, Z. Anorg. Allgem. Chem., 199, 321 (1931). . J. F. Elliott and M. Gleiser, Thermochemistry for Steel- making, Addison-Wesley Publishing 00., Cambridge, Mass. (1960). T. Ellis, I. M. Davidson and C. Bodsworth, Journal of the Iron and Steel Institute, 201, 582 (1963). Leslie M. Ferris and Mildred J. Bradley, JACS, 81, 1710 (1965). ' S. A. Golovenenko, I. A. Tomilin, I. Yu. Konnova, IZU Akad Nauk SSR Met (Russian Metallurgy), g, 72 (1973). S. R. de Groot, Thermodynamics pf Irreversible Processes, North Holland Publishing Co. (1951). R. Haase, Thermodynamics of Irreversible Processes, Addi- son—WesIey, Reading MA (1969). M. Hansen and K. Anderko, Constitution 9: Binary Alloys, McGraw-Hill, New York (1958). G. Herzberg, Molecular Spectra and Molecular Structure 1 Diatomic Molecules, Prentice Hall, Inc., New York (1939). p. l98f. G. Herzberg and K. N. Rao, J. Chem. Phys. 11, 1099 (1949). K. F. Herzfeld and W. Heitler, Z. Electrochemie, 31, 536 (1925). J. H. Hildebrand, JACS £1, 66 (1929). 245 J. H. Hildebrand,Proc. Nat. Acad. Sci. 13, 267 (1927). J. H. Hildebrand, Proc. Nat. Acad. Sci. 13, 167 (1927). F. H. Horne and T. G. Anderson, J. Chem. Phys. 53, 2321 (1970). W. Hume-Rothery, H. M. Irving, and J. R. P. Williams, Proc. Roy. Soc. A208, 431 (1951). JANAF Thermochemical Tables, Dow Chemical Company, Midland, MI (1971). L. Kaufman, Acta Met., 1, 575 (1959). L. Kaufman, Prog. Mat. Sci., 14, 2 (1969). L. Kaufman, Phase Stability in Metals and Alloys, (P. S. Rudman, J. Stringer, and R. I. Jaffer, eds. ), McGraw Hill, NY, 1967, p. l25f. L. Kaufman and H. Nesor, Ann. Rev. of Material Sci. (R. Huggin, ed.), Vol. 3, Annual Reviews, Palo Alto, CA (1973). L. Kaufman and H. Nesor, Treatise on Solid State Chemistry, N. B. Hannay, editor, Plenum Press, New York (1975), .2: 179- 232 L. Kaufman and H. Nesor, Computer Calculation of Phase Diagrams, Academic Press, New York (1970). E. Kenik and R. Carpenter, private communication, 0RNL (1977). J. M. Leitnaker, G. A. Potter, D. J. Bradley, J. 0. Franklin and W. R. Laing, Met. Trans., In Press. J. M. Leitnaker, unpublished results, 0RNL (1977). F. Kohler, Monat. Fur Chemie, 21, 738 (1960). G. N. Lewis and M. Randall, Thermodynamics, 2nd Ed.,Revised by K. Pitzer and L. Brewer, McGraw Hill, New York, 1961. 0. Kubaschewski, K. Geiger and K. Hack, Z. Metallkde., §§. 337 (1977). C. H. P. Lupis and J. F. Elliott, Acta Met., £1, 529 (1966). M. L. McGlasken, JPAL, g1, l (1970). [1163053 (I) QW'TJ'U'U 211315133194 246 Newmaun, H. Schenck, and W. Patterson, Giesserei., Tech.-Wiss. Beih. Giessereiw. Metallk. 23, 1217. 1959. A. Oriani, J. Phys. Chem. Solids 30, 339 (1969). K. Plyer, L. R. Blaine, and W. S. Connor, J. Optical Society of America, 45, 102 (1955). H. Rank, D. P. Eastman, B. S. Rao, and T. A. Wiggins, J. of the Optical Society of America, 51, 929 (1961). D. Richardson, J. Iron and Steel Inst. 175, 33 (1953). Scatchard, Chemical Reviews, 44, 7 (1949). Scatchard, Chemical Reviews, 8, 321 (1931). Scheil, T. Schmidt, and J. Wunning, Arch. Eisenhiitt, 12. 251 (1961). Schneck, M. Frohberg, and E. Jaspert, Arch. Eisenh., 12. 683 (1965). Schurmann, T. Schmidt, and H. Wagener, Giesserei.- Forsch, 16, 91 (1964). Shewmon, Acta Met. 8, 605 (1960). Shewmon, J. Chem. Phys. 29, 1032 (1958). Shewmon, TMS-AIME, 212, 642 (1958). L. Shriver, M. Wuttig, Acta Met. 20, l (1972). K. Sigworth and J. E. Elliott, Canadian Met. Quart. 1;. 455 (1974). K. Sigworth and J. E. Elliott, Canadian Met. Quart. C. Slater, J. Chem. Phys. 41, 3199 (1964). P. Smith, TMS—AIME, gig, 1224 (1966). P. Smith, TMS—AIME, g1g, 62 (1960). P. Smith, JACS, 1Q, 2724 (1948). P. Smith, JACS, 6g, 1163 (1946). 247 H. Smith, TMS-AIME, 209, 47 (1952). P. Snow and E. K. Rideal, Pro. Royal Soc., London, A125, 462 (1929) E. Spear and J. M. Leitnaker, J. Am. Ceram. Soc., 52 257 (1969). R. Stover and J. Wulff, TMS-AIME, 215, 127 (1959). G. Ulitchny and R. Gibala, Met. Trans., 4, 497 (1973). S. National Bureau of Standards, Selected Values of Chemical Thermodynamic Properties TN270- 3, 1968. Wada, H. Wada, J. F. Elliott and J. Chipman, Met. Trans. 5, 2865 (1972). Wada, H. Wada, J. Elliott, and J. Chipman, Met. Trans., g, 2199 (1971). . Wagner, Thermodynamics of A110 s, Addison-Wesley Press Inc. Cambridge, Mass (1952 . W. White, P. J. Oenny, and S. M. Irving, The Electron Microprobe, (T. D. McKinley, K. F. J. Heinrich, and D. B. Wittry, editors); J. Wiley and Sons, New York (1966) E. Wick and F. E. Block, Thermodynamic PrOperties of 65 Elements Their Oxides, Halides, Carbides, and Nitrides, Bureau of Mines Bulletin 605, 1963. I'. ll il."«. I 111‘! 31’- .Ill.l J1