. t. 13:“; $1233: S" .5. T. - ’3 r353 .;. . .. 4. Rama}. :1 :2 n A .2? (an . 713;! y I. .23....LM JR . It f!!! .J a\ I! lit}. I . . .IL.‘ manual: ,. zl. . in}! 5.25” a .3 . , .3... .i... .. . x , y . x . . . , . . .w .o . . . A , , V «webs. .n: THESIS |||l|||||||llllllllllilllllllllllllllllllllllllllllllllll|||l| 1293 01801 7453 LIBRARY Michigan State University This is to certify that the thesis entitled An Analysis of Global and Localized Creep Strains and Creep Properties of Non-Composite and Composite Lead—Free Solders at Room and Elevated Temperatures presented by Jeffrey Lee McDougall has been accepted towards fulfillment of the requirements for Master's Materials Science degree in Major professor Dan: December 16, 1998 0-7639 MSU is an Affirmative Action/Equal Opportunity Institution PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 1/98 alClRCJDateDoepGS-p.“ AN ANALYSIS OF GLOBAL AND LOCALIZED CREEP STRAINS AND CREEP PROPERTIES OF NON-COMPOSITE AND COMPOSITE LEAD-FREE SOLDERS AT ROOM AND ELEVATED TEMPERATURES By Jeffrey Lee McDougall A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Materials Science and Mechanics 1998 ABSTRACT AN ANALYSIS OF GLOBAL AND LOCALIZED CREEP STRAINS AND CREEP PROPERTIES OF NON-COMPOSITE AND COMPOSITE LEAD-FREE SOLDERS AT ROOM AND ELEVATED TEMPERATURES By Jeffrey Lee McDougall Due to the adoption of “green” manufacturing practices in Europe, as well as the need for solders with better creep resistance at higher temperatures, there has been new interest in developing lead-flee solders. Understanding and quantification of the effects of creep deformation is essential for lifetime prediction of electronic systems. Miniature single shear-lap creep specimens, similar in size to actual solder joints used in electronic packaging, have been developed to examine the effect of in-situ composite microstructures on the creep resistance and damage accumulation in lead-free solder joints. Average and localized strain measurements have been made optically, by monitoring the displacement of a scratch across the thickness of the solder joint. By analysis of the local strains, criteria for damage initiation are expressed in terms of the onset of tertiary creep, which starts sooner in some regions of the microstructure than others, leading to heterogeneous strain in the joint. The in-situ composite solders result in a more homogeneous strain evolution, which causes an improved creep resistance at lower temperatures, and much greater uniform strain at higher strain-rates before a critical damage level is reached. This work is dedicated with love and gratitude to mom, Wendy, dad., and of course to grandma. Also, I would like to dedicate this to the memory of Gramma and Grampa and my great grandmother, Gram, my departed and current best friends, Casey and Drexyl, the Glennie Chapter of the MacDougall Clan for providing me with Scottish blood, Rico for being Rico, Tommy - who has shown me that one man’s ego can never be too large, Curtis for introducing me to Atlanta and showing me that space and time are meaningless in the realm of friendship, Joel - for his Scotch and song, Don - my eternal Glennie Buddy, Wade - for helping me realize a computer’s true purpose, Cousin Ed and all the boys up at Alcona Park, and Edna & Jeff Felmlee for making this all possible. Finally, I dedicate this to Sylvann for mending my broken faith and heart. iii ACKNOWLEDGEMENTS I would like to express sincere gratitude to Dr. Thomas R. Bieler for his guidance, support, friendship, and belief in my capabilities. Sincere thanks is also directed towards Dr. Lucas and Dr. Subramanian. I would also like to thank Dr. Case for placing in me the desire to become a better student. Finally, I want to express sincere appreciation for my colleagues and friends Sunglak Choi and Alan Gibson for their contributions, both academic and cultural. I would also like to acknowledge the Composite Materials and Structure Center for their financial contributions which made this research possible, Dong Yi Seo for the use of his Macintosh computer, Kurt Neimeyer for his contributions to the miniature creep frame, and my family and friends for their financial support and encouragement throughout my academic career. iv TABLE OF CONTENTS LIST OF TABLES ........................................................................................................... viii LIST OF FIGURES ............................................................................................................ ix CHAPTER 1: INTRODUCTION .................................................................................... 1 CHAPTER 2: REVIEW OF RELATED LITERATURE ............................................. 3 2.1 SOLDER OVERVIEW ............................................................................................ 3 2.1.1 The History of Soldering .................................................................................... 3 2.1.2 Solders and Soldering ........................................................................................ 5 2.2 CURRENT SOLDERING ISSUES .......................................................................... 7 2.3 MICROSTRUCTURE TAILORING ...................................................................... 11 2.4 CREEP OVERVIEW .............................................................................................. 13 2.4.1 History of Creep ............................................................................................... 13 2.4.2 Stress and Strain ............................................................................................... 14 2.4.3 Homologous Temperature ................................................................................ 15 2.4.4 Creep Testing ................................................................................................... 16 2.4.5 The Creep Curve .............................................................................................. 17 2.4.5.1 Secondary Creep-Rate ............................................................................... 20 2.4.5.2 Scattering of Data ..................................................................................... 22 2.4.6 Creep on an Atomic Scale ................................................................................ 22 2.4.7 The Role of Dislocations in Deformation ........................................................ 23 2.4.8 Steady-State Creep-Rate .................................................................................. 24 2.4.9 Secondary Creep-Rate Normalization .............................................................. 26 2.5 CREEP MECHANISMS ........................................................................................ 27 2.5.1 Dislocation Controlled Creep Mechanisms ..................................................... 27 2.5.2 Diffusion-Controlled Creep Mechanisms ........................................................ 28 2.6 CREEP-FATIGUE INTERACTIONS .................................................................... 30 2.6.1 Stress Relaxation .............................................................................................. 32 2.6.2 Fatigue .............................................................................................................. 33 2.6.3 Fatigue of Solder Joints .................................................................................... 35 2.7 CREEP IN SOLDER JOINTS ................................................................................ 40 CHAPTER 3: EXPERIMENTAL PROCEDURE ....................................................... 41 3.1 SAMPLE PREPARATION .................................................................................... 41 3.1.1 Joint Fabrication ............................................................................................... 41 3.1.2 Solder Microstructures ..................................................................................... 43 3.1.3 Polishing .......................................................................................................... 43 3.2 CREEP TESTING .................................................................................................. 47 3.2.1 Miniature Creep Frame .................................................................................... 47 3.2.2 Miniature Electrical Resistance Furnace .......................................................... 47 3.2.3 Deformation of Solder Joints ........................................................................... 49 3.3 DATA EXTRACTION TECHNIQUE ................................................................... 54 3.3.1 Image Enhancement ......................................................................................... 54 3.3.2 Data Extraction ................................................................................................ 55 3.3.3 Reproducibility of Data Extraction .................................................................. 58 3.4 DATA PROCESSING ............................................................................................ 59 3.4.1 Global Creep Strain Measurement ................................................................... 59 3.4.2 Localized Creep Strain Measurement .............................................................. 61 3.4.3 Strain-Rate, Position, and Time ....................................................................... 63 3.5 SEM OBSERVATIONS ......................................................................................... 65 3.5.1 Determination of Stress .................................................................................... 65 3.5.2 Initiation of Fracture ........................................................................................ 65 3.5.3 Location of Fracture ......................................................................................... 66 CHAPTER 4: RESULTS ............................................................................................... 67 4.1 COMPARISON OF EXPERIMENTAL CREEP DATA TO DATA FOUND IN THE LITERATURE ............................................................................ 67 4.1.1 Room Temperature Composite Data ............................................................... 69 4.1.2 Room Temperature Non-Composite Data ....................................................... 69 4.1.3 85°C Data ......................................................................................................... 70 4.1.4 125°C Data ....................................................................................................... 70 4.2 GLOBAL VS. LOCALIZED CREEP STRAIN ..................................................... 71 4.2.] Room Temperature Composite Creep Data ..................................................... 71 4.2.2 Room Temperature Non-Composite Creep Data ............................................. 73 4.2.3 85°C Composite Creep Data ............................................................................ 75 4.2.4 85°C Non-Composite Creep Data .................................................................... 75 4.2.5 125°C Composite Creep Data .......................................................................... 78 4.2.6 125°C Non-Composite Creep Data .................................................................. 80 4.3 POSITION OF MAXIMUM LOCALIZED CREEP STRAIN ............................... 82 4.4 FRACTURE SURFACES ...................................................................................... 83 4.5 RESULTS OF LOCATION OF F RACTURE EXAMINATION ........................... 86 4.6 RESULTS OF CRACK INITIATION TESTS ....................................................... 87 CHAPTER 5: DISCUSSION ......................................................................................... 90 5.1 CREEP BEHAVIOR ............................................................................................... 90 5.2 HOMOGENEITY OF DEFORMATION ............................................................... 91 CHAPTER 6: CONCLUSIONS .................................................................................... 96 APPENDIX A: CREEP MECHANISMS ......................................................................... 98 APPENDIX B: DATA REDUCTION ............................................................................ 100 vi APPENDIX C: DISPLACEMENT, STRAIN, AND STRAIN-RATE HISTORIES ..... 105 APPENDIX D: GLOBAL SECONDARY STRAIN-RATE DATA .............................. 125 APPENDIX E: FRACTURE SURFACES AND PROFILES ........................................ 136 APPENDIX F: SOLDER JOINT CREEP DATA .......................................................... 154 REFERENCES ................................................................................................................ 157 vii LIST OF TABLES Table 1. Creep Mechanisms .............................................................................................. 99 Table 2. Scratch Displacement Data. .............................................................................. 102 Table 3. Normalized Position and Modified Scratch Displacement Data ...................... 104 Table 4. Solder Joint Creep Data. ................................................................................... 155 viii LIST OF FIGURES Figure l. The Model Shape of a Typical Creep Curve. .................................................... 19 Figure 2. The Process of Stress Relaxation ....................................................................... 36 Figure 3. The Process of Thenno-Mechanical Fatigue ..................................................... 38 Figure 4. Differences Between Normal (a) and Accelerated (b) Thermo-Mechanical Fatigue Testing Adapted From Darveaux et al. ................................................ 39 Figure 5a. Assembly of the Solder Joint. .......................................................................... 42 Figure 5b. Finished Solder Joint with Magnified View of Joint Area. ............................. 42 Figure 6a. Eutectic 96.SSn-3.5Ag Non-Composite Solder Micrograph. ......................... 44 Figure 6b. Eutectic 96.58n-3.5Ag Composite Solder Micrograph. ................................. 44 Figure 7. Specimen Grip and Polishing Bar. .................................................................... 45 Figure 8. Development of the Miniature Creep Frame. .................................................... 48 Figure 9. Miniature Electrical Resistance Furnace. .......................................................... 50 Figure 10. Creep Frame Set Up for Elevated Temperature Testing .................................. 51 Figure 1]. Displacement of Scratch Due to Creep Shown as: (a) Raw Captured Image in Negative Format. ................................................. 53 (b) Enhanced Image for use in Datathief ........................................................ 53 Figure 12. Normalized Position Across the Thickness of a Solder Joint ........................... 56 Figure 13. Screen Capture of Data Extraction Using Datathief. ....................................... 56 Figure 14. Scratch Displacement History. ........................................................................ 60 Figure 15. Difference Between Global and Localized Deformation ................................. 60 Figure 16. Cubic Polynomial Curve Fit. ........................................................................... 62 ix Figure 17. Localized and Global Creep Strain. ................................................................. 62 Figure 18a. Variation of Strain-Rate to Time and Normalized Position for a Composite Solder Creep Tested at 25°C ........................................................................... 64 Figure 18b. Variation of Strain-Rate to Time and Normalized Position for a Non- Composite Solder Creep Tested at 85°C. ....................................................... 64 Figure 19. Comparison of Secondary Creep-Rate Data. ................................................... 68 Figure 20. Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 25°C. .......................................................... 72 Figure 21. Displacement (a), Creep Strain (b), and Strain-Rate (c) Histories for Non- Composite Solder at 25°C. ............................................................................... 74 Figure 22. Displacement (a), Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 85°C. ............................................................................... 76 Figure 23. Displacement (a), Creep Strain (b), and Strain-Rate (c) Histories for Non- Composite Solder at 85°C. ............................................................................... 77 Figure 24. Displacement (a), Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 85°C. ............................................................................... 79 Figure 25. Displacement (a), Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 85°C. ............................................................................... 81 Figure 26a. Non-Composite Fracture Surface, T = 25°C, I = 21.7 MPa. (Sample 038)...84 Figure 26b. Non-Composite Fracture Surface, T = 25°C, I = 23.5 MPa. (Sample 040). .84 Figure 26c. Composite Fracture Surface, T = 25°C, I = 32.2 MPa. (Sample 052). .......... 85 Figure 26d. Composite Fracture Surface, T = 125°C, 1.’ = 9.1 MPa. (Sample 036) ........... 85 Figure 27. Formation of Circular Areas Under High Stress Conditions: (a) Formation of Circular Areas. .................................................................... 88 (b) ‘Pop’ Out Upon Further Deformation. ..................................................... 88 (c) Collapse of Circular Area. ........................................................................ 88 Figure 28. Solder Joint Micrographs Before (a) and Afier (b) Further Deformation. ...... 89 Figure 29. Onset of Local Tertiary Creep vs. Global Secondary Creep-Rate. .................. 93 Figure 30. Figure 31. Figure 32. Figure 33. Figure 34. Figure 35. Figure 36. Figure 37. Figure 38. Figure 39. Figure 40. Figure 41. Figure 42. Figure 43. Figure 44. Figure 45. Strain Homogeneity Ratio vs. Global Secondary Creep-Rate. ........................ 93 Homogeneity of Deformation vs. Average Percent Porosity. .......................... 94 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 25°C. ........................................................ 106 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 25°C. ........................................................ 107 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 25°C. ........................................................ 108 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 25°C ................................................. 109 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 25°C ................................................. l 10 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 25°C ................................................. 1 l 1 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 25°C ................................................. 1 12 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 25°C ................................................. l 13 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 85°C. ........................................................ 114 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 85°C. ........................................................ 115 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 85°C ................................................. 1 l6 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 85°C ................................................. l 17 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 85°C ................................................. 118 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 125°C. ...................................................... 119 xi Figure 46. Figure 47. Figure 48. Figure 49. Figure 50. Figure 51. Figure 52. Figure 53. Figure 54. Figure 55. Figure 56. Figure 57. Figure 58. Figure 59. Figure 60. Figure 61. Figure 62. Figure 63. Figure 64. Figure 65. Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 125°C. ...................................................... 120 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 125°C. ...................................................... 121 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 125°C. ...................................................... 122 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 125°C ............................................... 123 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 125°C. .............................................. 124 Global Secondary Strain Rate for 25°C, 31.1 MPa, Composite Solder ......... 126 Global Secondary Strain Rate for 25°C, 28.9 MPa, Composite Solder ......... 126 Global Secondary Strain Rate for 25°C, 25.8 MPa, Composite Solder ......... 127 Global Secondary Strain Rate for 25°C, 21.9 MPa, Non-Composite Solder 127 Global Secondary Strain Rate for 25°C, 19.9 MPa, Non-Composite Solder 128 Global Secondary Strain Rate for 25°C, 19.8 MPa, Non-Composite Solder 128 Global Secondary Strain Rate for 25°C, 10.5 MPa, Non-Composite Solder 129 Global Secondary Strain Rate for 25°C, 11 MPa, Non—Composite Solder ...129 Global Secondary Strain Rate for 85°C, 10.8 MPa, Composite Solder ......... 130 Global Secondary Strain Rate for 85°C, 10.4 MPa, Composite Solder ......... 130 Global Secondary Strain Rate for 85°C, 12.3 MPa, Non-Composite Solder 131 Global Secondary Strain Rate for 85°C, 17.3 MPa, Non-Composite Solder 131 Global Secondary Strain Rate for 85°C, 16.3 MPa, Non-Composite Solder 132 Global Secondary Strain Rate for 125°C, 9.4 MPa, Composite Solder ......... 132 Global Secondary Strain Rate for 125°C, 9.9 MPa, Composite Solder ......... 133 xii Figure 66 Figure 67 Figure 68 Figure 69 Figure 70. Figure 71. Figure 72. Figure 73. Figure 74. Figure 75. Figure 76. Figure 77. Figure 78. Figure 79. Figure 80. Figure 81. Figure 82. Figure 83. Figure 84. Figure 85. Figure 86. . Global Secondary Strain Rate for 125°C, 11 MPa, Composite Solder .......... . Global Secondary Strain Rate for 125°C, 8.1 MPa, Composite Solder ......... . Global Secondary Strain Rate for 125°C, 9.7 MPa, Non-Composite Solder . Global Secondary Strain Rate for 125°C, 31.1 MPa, Composite Solder ....... Fracture Surfaces and Profile for 25°C, 31.1 MPa, Composite Solder ........... Fracture Surfaces and Profile for 25°C, 28.9 MPa, Composite Solder ........... Fracture Surfaces and Profile for 25°C, 21.9 MPa, Non-Composite Solder. Fracture Surfaces and Profile for 25°C, 19.9 MPa, Non-Composite Solder. Fracture Surfaces and Profile for 25°C, 19.8 MPa, Non-Composite Solder. Fracture Surfaces and Profile for 25°C, 10.5 MPa, Non-Composite Solder. Fracture Surfaces and Profile for 85°C, 10.8 MPa, Composite Solder ........... Fracture Surfaces and Profile for 85°C, 10.4 MPa, Composite Solder ........... Fracture Surfaces and Profile for 85°C, 12.3 MPa, Non-Composite Solder. Fracture Surfaces and Profile for 85°C, 17.3 MPa, Non-Composite Solder. Fracture Surfaces and Profile for 85°C, 16.3 MPa, Non-Composite Solder. Fracture Surfaces and Profile for 125°C, 9.4 MPa, Composite Solder ........... Fracture Surfaces and Profile for 125°C, 9.9 MPa, Composite Solder ........... Fracture Surfaces and Profile for 125°C, 11 MPa, Composite Solder ............ Fracture Surfaces and Profile for 125°C, 8.1 MPa, Composite Solder ........... Fracture Surfaces and Profile for 125°C, 9.7 MPa, Non-Composite Solder. . xiii Fracture Surfaces and Profile for 125°C, 10.9 MPa, Non-Composite Solder. 133 134 134 135 137 138 .139 .140 .141 .142 143 144 .145 .146 .147 148 149 150 151 152 153 CHAPTER 1 INTRODUCTION As electronic components continue to develop functionally, and less expensive means of manufacturing are developed, the world becomes continually more dependent upon electronic technology. However, the use of lead bearing solders is problematic as countries strive to become more health and environmentally conscious. Outlawing lead bearing solders would drastically alter the manufacture of electronic components since eutectic Sn-Pb is the most widely utilized of all solders. Today’s electronic components are more powerful, smaller, and relatively inexpensive in respect to their functionality. Although it has been in existence for thousands of years, soldering technology still represents the vast majority of reliable electrical connections. Unfortunately, as the functionality and miniaturization of electronic components continues, the operational temperatures and stresses increase. Use of soldered components in relatively high temperature environments such as under-the— hood automotive applications also makes greater demands upon soldering technology. Issues such as high temperature creep become relevant, considering that room temperature is over half the melting point (183°C) of eutectic tin-lead solder. As a result, alternative solders which provide higher performance must be investigated. In order to combat the detrimental effects of higher operational temperatures, solders with greater melting temperatures and creep resistance must be developed. Currently, there has been much interest in the use and development of alternative solders. However, much of the experimentation is being performed on bulk solder specimens possessing microstructures unlike the miniature, rapidly solidified (relatively small grain size) solder joints encountered in industrial practices (See Section 2.2, page 9). There have been few investigations utilizing miniature solder joint specimens. Investigations involving the creep characteristics of solders have dealt with creep in the classic sense, where the overall or global creep of the specimen was measured. Global creep data yields results seen in a typical or classic creep curve. Along with classic global creep, this investigation also considered creep from a localized perspective, i.e., instead of an measuring the average or global creep strain, creep strains were measured at different locations across the thickness of the solder joint. The scope of this research involves examination of creep resistance of miniature industrial-like solder joints produced from eutectic tin-silver solder with and without added interrnetallics as reinforcements. Creep characteristics of these solders will be examined at room and elevated temperatures. New methods of plotting the localized creep strain-rate were explored. An attempt to determine a criterion for the onset of tertiary creep deformation is made. CHAPTER 2 REVIEW OF RELATED LITERATURE 2.1 S OLDER OVER WE W 2.1.1 The History of Soldering The history of soldering dates back to over 6000 years ago [1]. Artisans from Mesopotamia first used solders in the form of gold-based, hard alloys to join together copper plates. Egypt was the next region to utilize soldering. Artifacts dating back to 3600 BC were discovered there, as well as in Greece and other Mediterranean areas, roughly 1000 years later. The most sophisticated example of early soldering comes from the Bronze Age where the Etruscans produced a gold pendent dating back to 2500 BC, in which three different types of solders were used within a 5mm proximity. This discovery was quite surprising, due to the fact that these three different solder alloys had varying melting temperatures. This is the earliest known example of a soldering process referred to as ‘step joining’. Evidence of soldering in the western hemisphere dates back to only around the year 10 AD. Remnants of the La Tolita natives of South America also indicated the use of gold-based solders. Currently, it is virtually impossible to determine whether or not these ancient civilizations were familiar with the use of soft tin-lead solders that are so prevalent today. Tin decomposes over time when exposed to tannic acid, a substance found in such organic entities as tree bark. As a result, soldering alloys which may have contained lead are indistinguishable from lead-free solders. Geographic evidence supports the notion that the Mediterranean region was not familiar with Sn—Pb solders since tin was not readily available. However, rich deposits of tin are found in northern Europe where soft soldering practices were developed. Historical evidence suggests that sofi soldering was developed by the Celts and Gauls from northern and central Europe around 1900 BC. Their exceptional use of soft soldering techniques were recorded by the Romans. Cooking utensils and metals tools could easily be manufactured or repaired over a small wood fueled fire, due to the low melting temperature of the tin-lead alloys, which were also high enough to withstand temperatures necessary for the preparation of food or hot water. Crude examples of soft soldering dating back to 1350 BC were discovered upon the unearthing of King Tut’s tomb. Evidence also shows that by 350 BC, the Greeks were using tin-lead solders to seal pumps and other devices produced from bronze. Tin-lead solders were also readily employed by the Romans to join the lead sheets which lined the interior of aqueducts. The application of soldering technology was greatly broadened during the Industrial revolution [1]. Due to their low cost, availability, and ease of use, tin-lead solders have prevailed through the twentieth century. With the development of portable heat sources, the use of soldering in such areas as plumbing, food container sealants, and the automobile industry vastly increased. It was not until the Second World War, when the supply of tin suddenly decreased, that alternative solders were given significant attention. Lowering the tin content as well as developing tin free solders was the major concern, but problems of higher melting temperatures and decreased fluidity were identified [2]. Today soldering is increasingly relied upon as the primary method of joining electronic components including resistors, capacitors, and packaged integrated circuits to ceramic substrates and printed wiring boards [3]. 2.1.2 Solders and Soldering Solders can be classified as being either soft alloys or hard alloys; 350°C is the melting temperature that distinguishes the two categories. Solders melting below 350°C are considered to be soft solders and they typically contain both lead and tin. Other soft solders can be composed of such elements as bismuth, indium, and cadmium. Hard solders have melting temperatures which are greater than 350°C and they contain metals such as gold, aluminum, zinc, and silicon [1]. The process of joining two metals together via a filler metal with a lower melting temperature than the materials being joined is known as soldering [1]. The soldering process simply relies on the wetting of the substrate materials by the solder for the formation of the bond [4]. Soldering is not the only means of joining two metallic materials. Two other similar methods include brazing and welding. Brazing is a similar metallurgical joining process which occurs at a temperature greater than 425°C. Brazing also relies on the wetting of the substrate materials, but some diffusion occurs that improves the bond strength [4]. When melting of the substrate materials in the region near the bond occurs, this is known as welding. A filler material can be used, but it is not always required in this process. Diffusion of the base metals is the means for bond formation [4]. Welding is a permanent joining technique, whereas soldering and brazing processes can be reversed with the application of sufficient heat. Often it is necessary to incorporate a flux into the soldering process. Fluxes serve a variety of purposes. They are used to clean the solder and substrate surfaces to be joined, as well as a protective coating before and during the soldering operation [5]. Fluxes were developed via the metal coatings industry where a clean surface was achieved prior to coating by immersing the substrate in an acid bath [6]. Fluxes are generally comprised of organic chemicals and organic salts. A natural resin from pine tress know as rosin can also be used. Rosin has the added benefit of not reacting until elevated temperatures are achieved [5]. The nomenclature for solders is based upon the alloy composition. The solder is described by its composition in weight percent from the most abundant to the least abundant element. For example, eutectic Sn-Pb solder (628n-38Pb), the standard soft solder, has a composition of 62 weight % tin with lead as the balance. Solder composition is quite a diverse matter. There are a wide variety of solders available which allows for a wide variety of specialized uses. As a result, the soldering process is not a simple matter of simply surface cleaning along with the addition of heat. There are many soldering techniques available, and the most prevalent issues concern types of solders, fluxes, base materials and their finishes, cleaning agents, and the design of the solder joint [3]. 2.2 CURRENT SOLDERING ISSUES Today, issues concerning solders and soldering are of paramount importance. Electronic components are being utilized in an ever increasing number of applications. The continual miniaturization of more powerful electronic components, makes the soldering process increasingly difficult. Along with miniaturization, higher temperatures are generated in certain applications. Also there is the continual concern of legislation or marketing pressures that could regulate or ban lead, which would dramatically affect the electronics industry in the United States. Many lead bearing products have already been outlawed due to noted effects on health as a result of prolonged exposure. Some of these products include paint and crayons, and the use of lead bearing solders in plumbing supplies were restricted in 1986. Internal combustion engines which could operate on unleaded gasoline were produced in the early 1970’s. Means of limiting human exposure to lead are continually being researched [1]. Even though solders comprise about 1% of the lead used in the US, there have been several attempts made by Congress, including a very recent one, to further reduce the use of lead, or tax its use severely. Lead bearing solders utilized in electronic components have yet to come under restriction in the US, although a number of European nations have already banned the use of lead bearing solders. This is a very relevant issue for American companies desiring to do business with European nations. Greater emphasis on the integrity of solder joints has resulted from the heightened functionality of electronic components [3]. Since the world is relying more and more on electronic components, solder joint failures must be kept to a minimum and designed in the most efficient manner. Many of the failures occurring in electronic systems are due to the failure of solder joints, not the components themselves [1]. In the earlier part of the twentieth century, as soldering became the most reliable means of electrical connections, the solder joint had only a single purpose which was just to maintain electrical continuity between components [1]. With today’s silicon chip technology, solder joints are also expected to provide structural support as well. The continual decrease in component size along with the need for solder joints to provide both electrical and structural support results in an increasing demand for stronger and more creep resistant soldering alloys [1]. It is well known that eutectic Sn-Pb solder coarsens after aging, leading to shortened life to failure. Thus we need to find a better alloy system that not only has no lead content but also possesses higher thermal fatigue resistance [7]. Current tin-lead solders lack shear strength, resistance to creep, and resistance to thermo-mechanical fatigue [8]. Eutectic Sn-Ag solder has a melting point of 221°C (compared to 183°C for eutectic Sn- Pb) so it may be suitable for higher temperature applications such as under-the-hood electronics in automobiles [9]. In most electronic systems, differential thermal expansion induces stresses resulting in substantial thermal cycling deformation of solder joint, which leads to eventual failure [10]. Soft solders are commonly used for electrical interconnects between components in electronic equipment. The stresses induced in these low melting temperature materials resulting from thermal expansion mismatch between the joined materials eventually leads to cracking that can limit component reliability [10]. While fatigue deformation is a major concern for electronic solders, creep constitutes an important state of deformation since stress relaxation occurs after a temperature change. In realistic thermal cycles there is sufficient time for stress relaxation processes to occur and creep induced damage may result [10]. It has been suggested that solder creep behavior may be directly associated with the joint fatigue lifetime [11]. Due to difficulties in experimental procedures, much of the data published concerning solder microstructure and mechanical properties has been based upon bulk solder rather than actual solder joints. Disagreement has existed regarding the use of bulk solder data in the design and analysis of actual solder joints. The microstructures of bulk solders differ from those of solder joints in significant ways [9]: Bulk solder does not have Cu intermetallics that form as a result of interactions between the solder and the Cu substrate. Solder joints are significantly stronger than the bulk tensile specimens [11]. These intermetallics (whether in the solder or in the solder/base metal interface) reduce the joint ductility. Secondly, large bulk specimens generally cool more slowly than joints on real printed wire boards, greatly affecting the creep properties of the solders. The creep behavior of the solder alloy may in large part determine the joint fatigue lifetime and may be the best means by which to understand the fatigue results [11]. The Sn-Ag shear-lap specimens were significantly stronger than the bulk tensile specimens [11]. Comparatively, there have been fewer investigations utilizing miniature solder joint specimens. There is a need to develop new solders that have similar processing properties to tin- lead solders, but which are lead-free and have improved mechanical properties and microstructural stability [8]. Coarsening during aging is more difficult in eutectic Sn-Ag solder because of the thermally stable intermetallics, therefore its rate of coarsening is much slower then that in eutectic Sn-Pb [7]. This indicates that known phenomena about Sn-Pb solders is not indicative of phenomena in other solders. 10 2.3 MICROSTRUCTURE TAILORING The microstructure of a metal or alloy exposed to temperatures close to half of its melting point can become dynamic. Due to the processing history of a material, a variety of microstructures may result just for one particular composition. An investigation performed by Frost et al. indicated that creep behavior of solder alloys was continually influenced by microstructural changes occurring at elevated temperatures [12]. Experimental evidence has shown that Sn-Ag and Sn-Pb solder microstructures can undergo significant coarsening at temperatures similar to operational temperatures observed in under-the-hood automotive conditions [13, 14]. Eutectic Sn-Ag solder joints are characterized with a relatively fine microstructure with well dispersed Ag3Sn intermetallics [13] which is excellent for fatigue resistance, but for creep resistance a coarse grain size is generally desired. The well dispersed Agg, Sn intermetallics observed in eutectic Sn-Ag solder can show significant growth after only 100 hours at temperatures between 60 and 150°C. The naturally forming Cu68n5 intermetallics at the Cu substrate/solder interface have also been shown to coarsen [14]. The presence of these brittle intermetallics at the substrate/solder interface can effect solder joint reliability. Loss of mechanical strength especially under high loading conditions is associated with these interfacial intermetallics [1]. In order for Sn-Ag solders to considered for future use under these high temperature conditions, the microstructure needs to be stabilized. By introducing a second phase (dispersoids) into an existing microstructure, grain boundary mobility can be inhibited which would result in improved microstructural ll stability. In addition to inhibiting grain growth, a second phase may also cause a more homogeneous distribution of strains [13]. Strengthening of the solder alloy against creep and fatigue deformation can also result through the addition of a second phase [15]. The composite solder used in this investigation contained a dispersed phase of Cu68n5 intermetallics on the order of 20 volrune percent. Upon initial examination, the intentionally added intermetallics indicated no improvement in controlling Ag3$n particle coarsening, but the added Cu6$n5 particles appeared to be quite stable [13]. In an investigation on Bi-Sn eutectic solder alloy [15] coarsening was suppressed and creep resistance greatly enhanced throguh the addition of Fe particles A comparison of creep properties of aged and unaged composite solder as compared to that of non-composite eutectic Sn-Ag solder was performed by Choi et al. The composite solder demonstrated dramatic decreases in the strain-rate as compared to the non-composite solder. Strain-rates decreased by a factor of 100 for aged specimens to 1000 for the unaged specimens [10]. Microstructural observations showed no displacement of the added CuéSns particles for specimens with strain-rates on the order of 10'7 3". Displacement was apparently limited to the Sn matrix. Under similar loading conditions, the composite solder showed a decrease in creep strain as compared to the non~composite solder. As a result, less stress relaxation may occur forcing the failure process away from the solder and instead to the components it joins [13]. 12 2.4 CREEP OVER VIEW Materials under stress at elevated temperatures can undergo a phenomenon know as creep. This phenomenon of continuous deformation under a constant load can occur at all temperatures above absolute zero and is also observed in non-metals [16]. This time- dependent strain includes elastic, viscous, and plastic deformation, but usually only plastic deformation is measured [16]. Creep strain is probably the most important time- dependent damage accumulation factor affecting solder joint reliability [17]. Problems usually associated with creep occur only at high homologous temperatures [18]. For design purposes, components used under stress at elevated temperatures must be capable of resistance to distortion over the life span, as well as creep failure should not occur within the required operating life [18]. Processes of creep deformation at low temperatures, less than about fifty percent of the melting temperature, are believed to be related to non-diffusion controlled mechanisms, while creep occurring above this temperature are generally associated with diffusion controlled mechanisms [16]. However, there still exists a great deal of controversy concerning mechanisms of creep deformation. 2. 4.1 History of Creep The phenomenon of time-dependent strain, or creep, was discovered hundreds of years ago. One of the earliest examples involved churches from the medieval period. Many of these churches were adorned with steep pitched roofs composed of lead. Since the roofs were at such a sharp angle, the lead would stretch under its own weight in a manner similar to liquid running down an inclined plane. This resulted in cracked roofs that 13 needed repair [19]. Another example of creep was discovered in domestic water supplies consisting of lead pipes. Gradual sagging of the lead pipes would occur which was not much of a concern, until this caused cracking and eventual bursting of the pipes. The sagging was more pronounced in pipes which supplied the hot water. Cold water pipes had fewer problems [19]. In terms of significant recorded work on creep, there are several important dates [16]. In 1905 Philips demonstrated that the creep of metals at low stresses was possible. Andrade also supported this theory in 1910 and again in 1914, however serious consideration was not given to the creep phenomenon until 1919 when Chevenard performed experimental work demonstrating that creep was not exclusive to just soft metals. Other significant research was accomplished by Dickenson in 1922. Major work on the theoretical aspects of creep was done by Becker (1925, 1926) and Orowan in 1934 [16]. 2. 4.2 Stress and Strain Considering the general properties of a typical stress-strain curve, no plastic deformation will occur until the stress increases beyond the proportional limit of the material, where the onset of yielding occurs. However, if the material is held under a constant load below the yield stress at sufficiently high temperature conditions, plastic deformation can occur. This temperature and time-dependent deformation is known as creep and it can occur at stresses below the macroscopic yield strength by at least one order of magnitude or greater due to thermal fluctuations which aid the motion of dislocations [l 6]. l4 Shear forces can also induce creep deformation [16]. Shear strain causes the displacement, dx, of a finite element in the direction of the shearing force in respect to the thickness, yo, of the element, where y0 remains constant. The shear strain, y, is given dc by, yxy = g. Strain is related to stress through Hooke’s law, which can be stated in . 0' . . T . . terms of normal strain, a = E’ and srmple shear strain, y = E. The varrable E rs known as the elastic or Young’s modulus and G is the shear modulus. Both moduli are temperature sensitive, i.e. with increasing temperature the moduli decrease. Most creep tests assume constant load 2. 4.3 Homologous Temperature Experimental evidence has shown that creep is an extremely temperature sensitive phenomenon where local thermal agitation provides the additional energy (beyond mechanical forces) necessary to overcome barriers to creep deformation [16]. The extent to which a material will deform under constant stress depends greatly on the material’s melting point. For example, a specimen of nickel (Tmeu =1453°C) under stress at room temperature may not exhibit any signs of creep, whereas a specimen of eutectic tin-lead solder (6OSn-40Pb, Tmen =183°C) under similar proportional stress conditions can creep excessively. An important term when considering the phenomenon of creep is a material’s homologous temperature which is defined as the ratio of the applied absolute - - 7:1 to temperature to the material’s absolute meltrng pomt, THomologous= ”I / 1 . "16 l The higher a material’s homologous temperature, the more likely it is to undergo creep deformation. At room temperature, nickel has a homologous temperature of around 15 0.17, whereas 6OSn-40Pb is at a homologous temperature of 0.65, a value over half of the absolute melting point. Temperatures where creep becomes important for pure metals are above 0.4Tmcn. The significance of this temperature resides in the fact that atomic rearrangements of the crystal lattice are possible via diffusion. The homologous temperature is an important parameter in the profile of creep curves as well as the onset of creep deformation mechanisms [18]. 2.4.4 Creep Testing In order to characterize creep properties, it is necessary to measure changes in the gage length with time. No completely satisfactory method of accomplishing this problem has yet been designed. The most common test method employed is the application of constant load (tension, but creep tests can also be done in compression) under constant temperature. Specimens used for a typical creep test are generally cylindrical with threaded ends which thread directly into the creep machine. The gage lengths typically range from 10 to 60 mm. It is important to consider that when testing at severe temperatures where oxidation of the specimen can occur, the specimen needs to be enclosed in a vacuum or inert atmosphere [20]. Currently the most common method for measuring this change is by using an extensometer in direct contact with ridges machined onto the gage length of the specimen or on to enlarged specimen shoulders. The change in gage length between the ridges can then be monitored via displacement of the extensometer arms. ' Some of the more common displacement detecting devices which can be used with the extensometer include dial gages, linear variable differential transformers (LVDT), and differential 16 capacitance transducers. Another method involves monitoring the displacement of scribe lines on the surface of the specimen via a venrier telescope focused through a viewing slot in the fumace. This requires manual operation and the scribe lines often disappear or become indistinct if the specimen surface oxidizes. [20] Creep tests can take anywhere from several minutes to several years. In general, it is desirable to avoid accelerated testing and carry out tests in real time. However, in order to ensure that successive batches of components are being manufactured under the customer guidelines, or for theoretical means such as determining how added composites affect the creep characteristics of a given material, accelerated creep tests are commonly conducted [18]. It is possible to predict the creep rupture life through a relationship developed by Monkman and Grant. They found that the rupture life is inversely proportional to the minimum or secondary creep-rate, as. The relationship is given by, log t, + Clog a, = K, where C and K are constants for a given material, and t, is the rupture life. 2.4.5 The Creep Curve Accumulation of strain with respect to time (strain history) is the basic information resulting from deformation under constant temperature and stress. The total accumulated creep strain can be described with the equation, 8,0,“, = so + e, where 80 represents the instantaneous strain upon loading, and e is the time-dependent strain. The creep-rate, a = ‘18 dt , can be determined by measuring the slope at any point on the creep curve. The resulting shape of the strain history profile is associated with different stages of creep. 17 51 (I) The first stage in creep, Stage I, is also known as primary creep which is a relatively short lived phenomenon. This transient period of creep represents a region of decreasing creep-rate. This stage of creep can be characterized by strain hardening. The second stage which can be seen from Figure l is known as secondary creep, steady-state creep, or Stage II. This stage is characterized by a nearly constant creep-rate that results from a balance between strain hardening and recovery. The final stage of creep, Stage III, also known as tertiary creep occurs from a decrease in cross-section area of the specimen either because of necking or the formation of internal voids or cracks which can be associated with metallurgical changes such as coarsening of precipitate particles, recrystallization, or diffusional changes in the phases present [16]. Tertiary creep is associated with high stresses due to the decrease in cross- sectional area. Throughout the creep experiment, the load is held constant. The axial stress increases as the specimen elongates and decreases in cross-sectional area. The initial stress is usually the reported stress, not the instantaneous stress [21]. The theoretical shape of a typical creep curve is shown in Figure 1. After an almost instantaneous strain of the specimen, 8,, due to loading, there is a decrease in the rate of creep (Stage I) which transforms into a steady state process (Stage II), where there is little noticeable rate change. This creep-rate then rapidly increases (Stage 111) until the Specimen fails [21]. Individual creep tests do not always yield all of the three stages. The Shapes of the creep curves resulting from individual tests can vary greatly according to temperature, test duration, and the stress value. For example, tests performed at low Stresses have a tendency to yield a sigmoidal (S-shape) curves as well as curves with short or long range oscillations [16]. 18 A l frooture . t 1 anory Secondary Tertiary {<1—-——— 1—--—-—-~r> <,_-_ —~—- 11 > ' l (o i {ff/ff;j_E[/—/If”! g‘ 5 ”fr”... £3:— creep rote Q l - ‘7’ I | ' ’i 1 so -‘ l _.._v___ _ _ _ _ _- _ ___ll -__ _ ___l Time, 7‘ Adapted from Dieter, page 439. Figure l. The Model Shape of a Typical Creep Curve. 19 Temperature has a notable effect upon the shape of a creep curve. Under low homologous temperatures, the total resulting strains are quite low (< 1%) and the creep curve may only enter stages of primary and secondary creep. At homologous temperatures in the range of 0.3 or less, the creep strain varies linearly with log(time) when crept in tension [16]. Failure seldom occurs for tests performed at low homologous temperatures [18]. However, as the temperature is increased to a homologous value near 0.4, the profile of the creep curve deviates from this logarithmic form [18]. In contrast, during creep tests performed at high homologous temperatures, all three stages of creep, including tertiary, will be experienced. Once tertiary creep is underway, failure is inevitable. 2. 4. 5.1 Secondary Creep-Rate The minimum creep-rate is the average value of the creep-rate during Stage 11. Generally, the creep characteristics of a material are described by values obtained from the ‘secondary’ or ‘steady state’ creep-rate, 8'5 , resulting from normal high temperature creep tests. This eliminates the need for a wider variety of equations representing the creep properties of materials tested at high homologous temperatures [18]. The creep-rate, é , can be determined by measuring the slope of the curve, d8 hit in Stage 11. Even though it is simple in theory, the measurement of creep resistance is a tedious task which requires extensive laboratory equipment. Typical tests can run over a wide range of times, from a few minutes to over ten years [21]. A linear relationship between the secondary creep-rate and temperature can be obtained by plotting ln(és) vs. (l/T). This indicates that the secondary creep-rate increases exponentially with temperature [18]. Temperature dependence of this nature 20 can be seen in other processes such as diffusion. Creep can be classified as an Arrhenius function [18]. This results in the proportion, as oc exp— (QC / RT) , where QC is the creep activation energy and R is the universal gas constant. This leads to the general form of the empirical power law creep equation which is a, = A0" exp[ k?) [22], where a, is the strain-rate or creep-rate, A is a constant, 0' is the stress which has a measurable exponent given as n, Q; is the creep activation energy, k is Boltzmann’s constant, and T represents temperature on the absolute scale. When this form of the equation is used to represent the creep process, values for the activation energies of pure metals in a high temperature environment, have been measured to be similar to the activation energies for lattice self diffusion [18]. This indicates that the mechanism for creep at high temperatures would be controlled by diffusion. This fact also explains the transition between low and high temperature creep behavior, since the movement of atoms occurs readily near a temperature of 0.4Tm [18]. Another equation developed by Dorn, which mixes metallurgical theory and the empirical power law is referred to as a semi-empirical , DGb " b p , equation. This is equation is given by 83 = A—kT—(g) (E) [23], where b rs the Burger’s vector, D is the diffusivity and is given by the equation D = D0 exp(%—%) , d is D . . . the grain size, p is the grain size exponent. The term RT rs referred to as the mobrlrty. 21 2.4.5.2 Scattering of Data Scatter is a quality inherent in creep data. This is due partly to assumptions made during a creep test. For purposes of convenience, temperature and stress are assumed to be constant for the duration of a creep test for proper characterization of materials [20]. These substantial assumptions are rarely the actual conditions experienced in practice. Deviations from these assumptions will result in errors in the determination of creep strain, strain-rates, and rupture lives [20]. Errors in the axiality (tension or shear) of loading as well as the skill necessary in setting up the test itself along with typical errors can arise just as in any experiment from imprecise measurement of specimen dimensions or a variety of other factors [20]. 2. 4.6 Creep on an Atomic Scale In order to understand creep, it is necessary to consider a material on the atomic level. Typically, a metal contains many individual crystals which are comprised of a simple lattice structure of atoms held together via atomic bonding. If a solid is subjected to a tensile force, the crystal lattice will adjust itself to oppose the applied force and maintain equilibrium. If a sufficient force were to be imposed upon this polycrystalline material, resulting in plastic deformation, the deformation is caused mainly from the movement of dislocations. The deformation which initially occurs is elastic or reversible. Once the elastic limit has been bypassed, the stress imposed is sufficient enough to cause some of the dislocations to move across the crystals, resulting in permanent or plastic deformation. After some deformation occurs, strain will quickly cease, unless the applied 22 load is increased. Deformation is halted because forces are generated to prevent further deformation. This phenomenon is know as work hardening. 2. 4. 7 The Role of Dislocations in Deformation As dislocations move through the material as a result of deformation caused by a stress higher than roughly 10'5 of the shear modulus [24], they multiply rapidly. A consequence of this rapid multiplication is dislocation crossing and tangling, which impedes the further movement of other dislocations. Unless the load is increased or extra energy is supplied in the form of heat, further deformation will not occur. In creep deformation this can be associated with Stage I as defined previously. The stress necessary to cause continued deformation in a material is known as the flow stress, 00 [22]. The flow stress can be defined as a material’s resistance to shape change and is a function of temperature, strain, and strain-rate [20]. The flow stress, 00 = or + oi , consists of a thermal component, 6]“, and, an athermal component associated with the dislocations inside the material, 6 i [22]. Recovery can be understood in terms of atomic vibration. The higher the temperature, the greater the degree of vibration and the less time needed for recovery to occur. The creep-rate reaches a constant value when the rate of strain hardening and rate of recovery become equal. Since atoms have a greater chance of moving from one atomic position to another when their degree of vibration is high, the creep-rate increases with increasing temperature. Diffusion of atoms can occur due to existing vacancies in the lattice. At elevated temperatures, experimentally calculated activation energies for creep usually equal the activation energy for self-diffusion [22]. This diffusion tends to make 23 vacancies move into the dislocation cores so that the dislocations shift from one plane of the crystal to another. This phenomenon is referred to as dislocation climb [19]. During climb of dislocations, the dislocations can either move or annihilate other dislocations of opposite sign. Since the diffusion of vacancies either towards edge dislocations (positive climb) or away from edge dislocations (negative climb) is required for edge dislocations to climb, creep deformation resulting from climb should involve the activation energy for self diffusion [22]. In this manner, the extensive tangling of dislocations sorts itself out and many dislocations disappear in a process known as recovery [19]. A material undergoing creep deformation may not do so in a homogeneous manner. Changes in the arrangement of dislocations in crystals has been observed using transmission electron microscopy. Some regions are characterized by being relatively dislocation free, while other areas possess a high density of dislocation tangles and networks [20]. As a result of this dislocation density imbalance, the specimen’s state of internal stress is not uniform since each dislocation has an associated stress field that affects the movement of nearby dislocations [20]. The difference between the applied stress and these internal stresses is the driving force for the motion of dislocations. As a result, the inhomogeneous nature of the stress field can exhibit itself in some peculiar ways [20]. 2. 4.8 Steady-State Creep-Rate Confusion exists concerning nomenclature for the creep-rate obtained from Stage II of a high temperature creep test. The common misconception is to refer to the second stage of 24 creep as ‘steady-state creep’. The term ‘steady-state’ implies an equilibrium process with respect to dislocations and is reached when the rate of work hardening is balanced by the rate of recovery. This concept forms the basis for most dislocation creep theories [20]. During steady-state creep, there are as many dislocations being created as there are being destroyed and the lifetime of each dislocation contributes one increment of strain. During steady-state creep, the dislocation density and the arrangement of the dislocations remains essentially unchanged which results in a constant strain-rate [20]. In general, measurable secondary creep-rates are found at stress levels greater than about 10'5 of the shear modulus and at temperatures greater than 04me [24]. No changes in microstructure will occur under steady-state conditions. A constant strain-rate may occur during secondary creep, but there may be changes in the microstructure such as local recrystallization near particle interfaces, etc. The linear portion of secondary creep may be misleading. In order to demonstrate true linearity, the strain-rate needs to be plotted against strain. The minimum strain-rate, which is equivalent to the steady state or secondary creep-rate, represents a balancing condition between the hardening and softening processes in the material. If the microstructure is not in a state of equilibrium, steady-state creep is not occurring. In fact very few materials undergo true steady-state creep, such as some pure metals and simple alloys. There is little convincing experimental proof justifying the existence of a steady-state creep component [16]. Unless evidence indicates the microstructure is unchanging, creep occurring during Stage II is better referred to as secondary creep. 25 2. 4. 9 Secondary Creep-Rate Normalization The way in which the creep-rate changes in respect to temperature can be shown by the relationship, as = A0" exp— (QC / RT), which is known as ‘power-law creep’. Considering secondary creep behavior, different values of n and Q, will be associated with different regimes of stress and temperature. It is possible to condense all of the different temperature creep data into one set through a normalization process. When secondary creep behavior can be described with the power-law, if both sides of the power law equation are multiplied by the factor, exp(Qc /RT), the result is, é, exp(Qc / RT) = A0”. The left side of this modified power law equation is referred to as the ‘temperature compensated secondary creep-rate’ [18]. A similar equation can be used for shear stress deformation. This modified function allows each log éx/logo relationship obtained from different temperatures to be superimposed onto a single line. The stress is also normalized with respect to temperature, 0' /E, where ‘E’ is the value of Young’s Modulus at the appropriate creep temperature [18]. 26 2.5 C. The pr< im'olx'c' llrese sliding exponc 2.5.1 1 Disloc therme mecha roughl dislocz dislocz They compe rearrar Wt creep, dominz elevate allowir 2.5 CREEP MECHANISMS The process of creep has considerable theoretical controversy as to the exact mechanisms involved. There exists at least four different types of creep deformation mechanisms. These include dislocation glide, dislocation creep, diffusion creep, and grain boundary sliding. A chart in Appendix A summarizes the creep mechanisms, creep and grain size exponents represented by values of n and p respectively, and activation energies, Q. 2.5.1 Dislocation Controlled Creep Mechanisms Dislocation glide occurs when dislocations moving along slip planes overcome barriers thermally assisted mechanisms involving the diffusion of vacancies or interstitials. This mechanism occurs at relatively high stresses as compared to other creep mechanisms, roughly 10‘4 <%<10‘2. The creep-rate is established by the ease with which dislocations are impeded by obstacles including precipitates, solute atoms, and other dislocations [18]. The basis for dislocation creep was developed by Orowan and Bailey. They concluded that the steady state creep-rate represents a balance between the competing factors of the strain hardening rate and the rate of thermal recovery by the rearrangement and annihilation of dislocations. Weertmann was responsible for the development of the earliest models of dislocation Creep. The models were based on a mechanism in which dislocation climb is the dominating factor. If an obstacle inhibits the movement of a gliding dislocation at elevated temperatures, a small amount of climb may allow it to overcome the obstacle, allowing it to glide to the next set of obstacles and the process is repeated. Almost all of 27 the strain is produced by the glide, however, the climb step controls the rate. Since diffusion of vacancies or interstitials is necessary for dislocation climb, the rate limiting factor is atomic diffusion. This model predicts an equation for the rate of creep where the stress exponent equals three, however, creep experiments with a range of metals show that the exponent on stress varies from 3 to 8, with a value of 5 being the average [21]. Thus for a range of medium to high stress levels at temperatures above half of the melting temperature, the steady state creep-rate is characterized by a power-law relation. When more than one mechanism is functioning, the creep behavior is controlled by the fastest mechanism. If there are mechanisms operating in sequence, the slowest mechanism will control creep deformation. 2. 5.2 Diffusion-Controlled Creep Mechanism Diffusional creep occurs under low stress, high temperature conditions. Diffirsional creep involves the flow of vacancies and interstitials through a crystal under the influence of . . 0' _ applied stress. Diffusron creep occurs for stress levels around G <10 4. Nabarro- Hening and Coble creep are included under this category [18]. When a polycrystalline metal creeps by the diffusion of atoms through the crystal lattice, the process is known as Nabarro-Herring creep [25]. When diffusion of atoms occurs along the grain boundaries, the process is known as Coble creep [26]. Although support for the idea of diffusional creep has been offered through experimentation and can be found throughout the literature [27], it is still a controversial topic. There is an ongoing argument concerning 28 the validity of Nabarro-Herring creep as well as a controversy over ‘denuded zones’ which were offered as the first immediate microstructural proof for diffusional creep [28]. 29 2.6 CREEP-FATIGUE INTERACTIONS Solder joints used for both mechanical and electrical contacts in electronic packaging devices normally do not fail as a result of a single high stress event [29]. Usually failure is a result of repeated or prolonged load applications that result in damage accumulation via creep-fatigue failure of the solder. The nature of the environment most solder joints are exposed to involves a variety of deleterious factors. Automotive applications provide an excellent test of the solder joint. Temperature fluctuations in the range of -40°C to 150°C are possible in automotive environments, along with mechanical vibrations in a range from 20 to 2000 Hz [30]. Fatigue can arise due to both the mechanical vibrations and fluctuations in temperature. As discussed previously, because of high homologous temperatures, fatigue behavior of solder joints is related to creep as well as the process of stress relaxation [31]. Solder joints can be considered as structural materials since they function not only as electrical interconnects, but also as mechanical bonds. In electronic devices, solder joints ofien constrain materials of differing coefficients of thermal expansion (CTE) that upon an increase in temperature due to internal heat dissipation from an electrical current or simply from environmental changes, will result in plastic strains [32]. Plastic strains are caused by CTE mismatch. Depending upon the size of the joint, even minute changes in temperature can have a drastic effect upon the solder and the component(s) it connects [32]. If the connected components have matched CTE’s, there is still a danger of plastic Strain due to differential rates of heating and cooling (temperature gradients) in localized regions [33]. When these electronic devices are repeatedly turned on and off, or 30 IE temperature cycles are created by the device itself, cyclic strains can result. Typically a sequential development [29] of local stressing, microcrack formation followed by crack initiation which may result from the accumulation of creep cavitation damage [31]. Crack propagation then ensues resulting in the failure of electrical solder joints. Evidence of creep fatigue interactions include fatigue straiations (fatigue) as well as intergranular cracking and voids on some of the grain surfaces (Creep) [34]. In order to improve upon the design and characteristics of solder joints, it is necessary to gain a better understanding of the mechanical and fatigue properties of solder alloys used in electronic components [35]. Currently, for a variety of solder alloys, fatigue data are available in the literature. However before applying this to the solder at hand, consideration must be given to solder composition, microstructure, design of the specimen, testing conditions, as well as the author’s definition of the fatigue. Plastic deformation occurs in many solder joint designs since the accumulated strains exceed the elastic limit of the solder [31]. If these factors are ignored, assessment of the fatigue life may be incorrect [3 5]. Another pertinent issue involves the duration of the actual fatigue lifetime. The actual lifetimes required for field use greatly exceed plausible time for testing. As a result, service reliability must be estimated through fatigue models [31]. In other words, reliable accelerated testing procedures must be developed. 31 2. 6.1 Stress Relaxation Most solders are characterized as being very compliant. When stress is applied to a solder joint, especially at elevated temperatures, the solder will yield in a manner to reduce the applied stress [36]. Stress relaxation and creep are two very closely related phenomena differing by which factor is held constant. Creep occurs when stress is held constant and the strain changes with time. However, when the strain is held constant, i.e., the material’s dimensions are unchanging, and the stress decreases either by way of static, dynamic, or creep conditions, the load it supports decreases with time. This process is known as stress relaxation [35, 16]. From a microstructural standpoint, the processes are virtually the same, and the profiles of strain-rate vs. stress for both processes are identical [36]. A further term, known as creep relaxation, refers to the time-dependent changes simultaneously occurring in both strain and stress [3 6]. A simplified model of stress relaxation was provided in a paper by Hall. When thermal stresses arise in a soldered assembly under conditions previously described, the components which are responsible for the applied stress on the solder are usually only stressed within their elastic range. A stiffness value, k, can be associated with this . dF . . assembly 1n the form k = —E [36]. If the solder yrelds (creeps) by a distance Ax, then the force on the joint will be decreased by a value of kAx, [36]. The equation, (a, —aC) Ay = T LAT — k—H [36], combines the shear strain, shear stress, and temperature components into one relation. The variable F is the force associated with each solder joint. The coefficients of thermal expansion of the solder and the component being 32 SO W It [In soldered are given by as and ac respectively. The height of the solder is given by H, and 2L is the in-plane dimension of the substrate. The effect described by the equation is easy to visualize, for example, if the temperature is held constant, the force will be related to the shear strain in a linear manner with the value of the slope being -kH [36]. Conditions will approach stress relaxation if the value of k is large or stiff, while low values of k imply creep processes [36]. Strain is associated with a compliant solder at high temperature conditions, as well as smaller forces since relaxation due to creep can occur almost instantly. On the other hand, stress is high with a stiff solder at low temperatures [36]. The accumulation of damage will be different for different parts of the thermal cycle since stress relaxation during the heating and cooling parts is not constant [31]. The incorporation of highly flexible leads with solder joints can greatly reduce creep strains developed as a result of cyclic loading conditions. A shearing of the solder joint will generally result when conditions do not allow for the use of flexible leads, such as in surface mounting [37]. However, an investigation by Ross et al. [17] indicates that the use of flexible leads can cause more damage than relief if certain criteria are not met. Findings indicated that lead stiffness must be at least 1000 times more flexible than the solder [17]. The process of stress relaxation is indicated in Figure 2. 2. 6.2 Fatigue Probably the most recognizable model for the prediction of fatigue life, Nf, is the Coffin- Manson relationship. The equation is based upon an empirical relation and can be found in a variety of forms. In order to better predict the fatigue life of solders, Coffin-Manson 33 We lead SITE to t et. lit: (13 fa type relations have been adjusted to account for temperature and frequency effects [3 8]. In order for this model to be utilized to predict the lifetimes of components with flexible leaded parts, the strain range must be the total plastic strain range. This must also include strain associated with creep deformation which must be determined accurately and added to the total plastic strain [17]. A8 -A A0 ‘8 The form included here, N f = [ g P] + {—0—} , was provided in a paper by Tien, f 0 et. al. The equation can be divided into two parts, low cycle and high cycle fatigue. The —A A6 first term, [jg—p] , describes low cycle fatigue which is controlled by the rate of crack ,- growth [38]. The plastic strain range Asp which is raised to the -A power is related to the -8 A0 damage stored in each cycle. The second term, {-0—} , is described as the high cycle fatigue component. During high cycle fatigue, the component spends most of its life initiating cracks and the damage stored in each cycle is best described by the stress range Ao raised to the -B power [38]. Fatigue life in the range of up 104 cycles is associated with low cycle fatigue, whereas fatigue life above 106 cycles is associated with high cycle fatigue [3 8]. The exponents A and B are respectively the low and high cycle fatigue Coffin-Manson variables. The fracture strain is given by 8f, while 00 is the stress associated with high cycle fatigue [38]. Vaynman, et al. found this model to be acceptable at strains above 102, when subjected Pb-Sn solders to isothermal fatigue. However, at strains below 102, the relationship failed due to strong creep fatigue interactions during cycling [35]. 34 2. 6.3 Fatigue of Solder Joints Fatigue in solder joints generally results from repeated thermal cycles caused by turning on and off electrical devices. The temperature fluctuations can induce cyclic strains as a result of thermal expansion mismatches between the solder and components, as previously mentioned, and result in the eventual failure of the component [32]. Prediction of the fatigue life of solder joints is a difficult task as a result of the time and temperature dependent characteristics of creep behavior [39]. A recent investigation into fatigue modeling has illustrated that creep is the dominating factor in the determination of fatigue life [40]. When considering fatigue life, there are a variety of relevant factors which must be considered including the range of temperature fluctuations, average temperature, dwell times, dwell temperature, rates of loading and unloading, joint geometry, solder microstructure, and the presence of alloying elements and intermetallics [39]. Fatigue damage is usually the result of dislocation pile-ups which cause stress concentrations. The complex microstructure of as-solidified eutectic solders as well as its instability at high temperatures make fatigue prediction models complicated [3 7]. Currently, no reliable predictive theory for thermal fatigue under creep conditions exists [37]. Recent attempts at modeling the complex interactions between creep and thermal cycling have indicated that the deformation and localized accumulations of strain produced during thermal cycling are not reversible [40]. This results fiom distortions occurring which change the stress distribution in the joint. Subsequent thermal cycles then have an altered impact on the state of stress and strain [30]. Experimental evidence indicates that strain is not reversible since continued thermal cycling results in 35 System is in Equilibrium 06, Component TReflow ins—0t. , or, PWB Component Reflow Soldered to Printed Wire Board 0t1>0t2>0t3 Removed from Reflow Furnace or, Compression <—————— <1— orI . - - 0% Tensron —————> PWB Shrinks Faster than Solder and Component Shear Stresses Result from CTE Mismatch Causing Shear Strains Resting Outside of Reflow Furnace . TRoom or, Compressron <~——— <»—Ot] . <— Olz Tensron —* Stress is Reduced as Shear Strain Increases Stress Relaxation Occurs Figure 2. The Process of Stress Relaxation 36 fi It in; C0 0C- deg permanent shape change. The process of thermo-mechanical fatigue is illustrated in Figure 3. Inaccuracies result when accelerated therrno-mechanical fatigue tests are performed. It has been shown [41] that under accelerated thermo-mechanical fatigue testing, the majority of plastic deformation occurs during loading of the specimen, whereas, in a test conducted under realistic heating and cooling rates, most of the plastic deformation occurs as a result of stress relaxation. During the cold part of the cycle, stress relaxation may not be as significant as during the warm part of the cycle since creep has a high dependence upon temperature [34]. As a result, the effects of creep are greatly ignored under accelerated test conditions. The differences are illustrated in Figure 4. 37 Low Temperature Dwell ' Figure 3. The Process of Thermo—Mechanical Fatigue 38 High Temperature Dwell High Temperature Dwell —j +1000 psi -0.005 \ + -1000 psi Low Temperature Dwell (a) - High Temperature Dwell +1000 psi —'— —,- -2000 psi Low Temperature Dwell (b) Figure 4. Differences Between Normal (a) and Accelerated (b) Thermo-Mechanical Fatigue Testing Adapted From Darveaux et al. 39 2.7 CREEP IN SOLDER JOINTS Recently, there has been much work done concerning the creep of solder. It has also been acknowledged in a variety of papers that there is a significant difference between testing easily fabricated bulk specimens of solder and testing actual solder joints. The discrepancies are related to a smaller grain size due to faster cooling rates during fabrication as well as the lack of substrate constraint in bulk specimens which eliminates the formation of intermetallics near and at the substrate/solder boundary. Both of these effects are believed to be responsible for the significant increase in strength of solder joints compared to bulk specimens. It has been noted [5] that solder containing silver (Ag) offer some of the highest creep resistances. This also results in reducing the strain observed during creep. Raeder, et al. [5] found in a study comparing Sn-Ag and Sn-Bi solder joints that at strain-rates greater than 10'4 s", Sn-Bi has a higher flow stress, and at strain—rates lower than 104 s", Sn-Ag solder has a higher flow stress. Raeder also noted significant creep differences between bulk solder and solder joint specimens. Studies on composite solders have also been conducted within the last several years [10, 15, 41]. In all instances, the addition of a dispersed phase significantly increased the creep resistance of the solder. It has also been noted that increased porosity may result from these additions. 40 CHAPTER 3 EXPERIMENTAL PROCEDURE 3. 1 SAMPLE PREPARA T I ON 3.1.1 Joint Fabrication The substrate material used in the fabrication of the solder joints consisted of copper dogbone halves, fabricated via electrical discharge machining (EDM). The substrate dimensions are shown in Figure 5a. In order to achieve a solder cross-sectional area of approximately 1 mm, a solder mask was applied with a fine brush, approximately 1 mm from the narrow end of the copper substrate. The freshly masked copper substrate halves were then heated at a temperature of 150 °C for one hour. Before the solder joints could be fabricated, the Cu substrates were chemically cleaned with a solution of 50% Nitric acid and 50% water. Once the Cu substrates were cleaned and dried, a fine drop of Alpha 200-L flux was applied to the narrow end of the dogbone shaped substrate. Half of the Cu substrates were then placed, flux side up, into a specially machined alruninum fixture shown in Figure 5a. Two pieces of solder foil, approximately 1mm2 in area and 30 microns in thickness, were sandwiched between the two Cu substrates as shown in Figure 5a. A thermocouple was connected to the aluminum fixture which was then placed directly onto a preheated hot plate. The fixture was heated to a temperature of 270°C on a hot plate and then quickly removed and placed on an aluminum chill block to simulate a cooling temperature history of industrial practice for reflowed solder joints. Once the aluminum fixture reached 41 £1.30 . mm P: 5111'?) I F1 1Sure g Solder Foil Alpha 200L Flux Solder Mask Thermocouple Wires 0.26 mm 58.0 mm Heated Aluminum Fixture Figure 5a. Assembly of the Solder Joint. Figure 5b. Finished Solder Joint with Magnified View of Joint Area. 42 ambient The ima mmSO Which \1 Nodm characte iis’lngu mama Visible z ambient temperature, the solder joints were removed and gently inspected by hand for integrity. The intact joints as seen in Figure 5b were then polished. The solder joints ranged in thickness from 50 to 100 pm. 3. 1.2 Solder Microstructures Two types of lead-free solder were used in this investigation. Solder joints were fabricated from eutectic tin-silver solder (96.58n—3.5Ag) with and without 20v% 5-8 pm CuGSns intermetallics which were added as reinforcements. Scanning electron microscope (SEM) micrographs of the two different solder microstructures are shown in Figure 6. The non-composite solder is characterized by a colony type microstructure. The colonies which are only slightly visible, are distinguished by minute Ag38n intermetallics visible in the grain boundaries. A layer of Cu6$n5 intermetallics which formed at the solder-substrate interface during the soldering operation is visible at the top of the micrograph. In the composite solder, Cu6$n5 intermetallics are clearly visible throughout the matrix. Naturally forming Cu68n5 intermetallics are also visible along the solder-substrate interface in the upper right comer. 3.1.3 Polishing During the polishing procedure, the solder joints were held with a special grip designed to properly support the specimen as shown in Figure 7. Wet/dry sandpaper (240 and 600 grit) was the first medium employed in polishing the solder joints. Next, a 0.5 micron polishing solution used with a standard metallurgical polishing cloth adhered to a narrow aluminum block was used to further polish the specimen. The specimen was then finished with a 0.05 micron SiOz colloidal suspension in the same manner. 43 Cu substrate “.1. C r168n5 lllltlmcldlllc’s I 96 556 3 5Ag s... ' 1' Solder Figure 6a. Eutectic 96.SSn-3.5Ag Non-Composite Solder Micrograph. Figure 6b. Eutectic 96.SSn-3.5Ag Composite Solder Micrograph. < 50.0 mm 7- , Polishing Cloth Aluminum Bar Figure 7. Specimen Grip and Polishing Bar. 45 In order to monitor displacement of the solder joints during creep testing, it was necessary to intentionally scratch the polished surface across both Cu substrates and the solder area. This is illustrated on page 53 with Figure 11. Several methods were used during this investigation. The first method was to rub the polished surface over a coarse polishing cloth. This method worked well for images taken with a standard optical microscope. However, the resulting scratches were too fine to be adequately monitored with the current image capturing system which incorporated a traveling microscope with a digital camera mounted on a tripod. A single-sided razor blade lightly pressed across the polished surface seemed to give the best results for this investigation. The scratch appears relatively large under a standard optical microscope, but when used in the current image capturing system, it provided sufficient results. 46 3.2 CREEP TESTING In order to subject the miniature solder joints to creep deformation at room and elevated temperatures, it was necessary to develop suitable testing equipment. The method chosen for free strain measurement required that the digital camera have an unobstructed view of the solder joint. This resulted in a variety of difficulties. Further problems were encountered when it was necessary to supply heat for elevated temperature creep tests. The devices and methods employed for this experiment underwent several stages of development until satisfactory testing methodology was achieved. 3. 2.1 Miniature Creep Frame The miniature creep frame underwent significant development throughout this investigation. The first design (Figure 8a) worked relatively well, but did not allow incorporation of the miniature furnace. It was simply an aluminum machined plate connected an I-beam. It consisted of one vertical slot in front which provided clearance for the suspended solder joint and another narrow slot traversing the top which allowed for the solder joint to be suspended with a piece of piano wire. The next version depicted in Figure 8b was developed to incorporate miniature furnace. 3. 2.2 Miniature Electrical Resistance Furnace A miniature electrical resistance furnace was constructed for use with the shear-lap soldered specimens. Nichrome, an alloy commonly used for electrical heating elements, due to its strong oxidation and electrical resistance, was used in wire form to create heating coils. In order to achieve even spacing of the coils, the Nichrome wire was 47 (b) Figure 8. Development of the Miniature Creep Frame. 48 wound around a 0.5 inch threaded bolt. The wound wire was then covered with a thin layer, approximately 1/4 inch, of kaowool, for thermal insulation. The assembly was then wound with fiber glass and coated with resin. In order to keep the fiber glass windings secure, three pieces of copper wire were fastened around the middle of the assembly and near the ends in a manner similar to the steel rings found on a wooden barrel. Excess copper wire was allowed on the two outer windings to be employed as electrical leads. The terminal ends of the Nichrome wire were fastened to the two outer copper wire windings. The furnace is shown in Figure 9. A variable voltage power supply was used to provide current for the miniature furnace. The temperature was monitored with a thermocouple. Teflon® coated type K thermocouple wires were fastened to the shear-lap specimen. The wires were twisted onto the copper substrate as near as possible to the solder joint without blocking the view of the camera. Coated wires were employed to ensure no electrical contact with the Nichrome wire coil. A heavier piece of copper wire was wrapped around the center of the fumace and secured onto the I-beam with a machine screw. The copper wire is strong enough to support the light weight furnace and also flexible so the fumace can be easily aligned with the hanging shear-lap specimen. See Figure 10. 3. 2.3 Deformation of Solder Joints The miniature solder joint was placed in the creep frame which was equipped with an electrical resistance heater (when necessary) and designed for dead load creep testing. The microscope and fiber optic lighting system were then positioned for optimrun resolution of details. The scratch must be clearly visible before testing begins, because as 49 Kaowool Nichrome Wire Coil Resin Coating Fiberglass Wrapping Figure 9. Miniature Electrical Resistance Furnace. 50 Support Solder Joint Length Extension Thermocouple Leads Electrical Resistance Furnace Figure 10. Creep Frame Set Up for Elevated Temperature Testing. 51 the creep test progresses, the scratch becomes less distinguishable (Figure 11). If the resistance firmace was used, the length of the solder joint was extended at the suspended end in order to allow exposure of the joint area to the microscope. This was done by sandwiching the end of the solder joint between two small lengths of metal and fastening the assembly with wire. Prior to loading, an image of the undisplaced scratch position was captured using a traveling microscope and a digital camera as a basis for comparison to subsequent images taken during the creep test. The specimen identification and load used were also recorded. Immediately afier loading the specimen with the desired weight, another image was captured. Care was taken to accurately record the time associated with each image. Depending upon the weight used, reliable experiments lasted anywhere from 10 minutes to more than three months, so this made accurate time measurements imperative for the short duration experiments. The frequency of image capturing depended upon the load used. Usually 10 to 30 data points are sufficient for generating an accurate creep curve. Variance in temperature as well as safety precautions during elevated temperature testing required the operator to remain present so that adjustments in temperature could be performed. The creep tests used for this investigation were performed at 25°C, 85°C, and 125°C with temperature fluctuations less than 7 degrees throughout the duration of each elevated temperature experiment. Both longer duration (low stress) and short duration (high stress) creep tests were performed at room temperature, but only short duration tests were performed for the elevated temperature creep. The short duration or accelerated tests were conducted in a range from 30 minutes to approximately 24 hours, while the long duration creep tests usually lasted for 2 to 3 months. 52 '11 Before Creep (a) (b) Figure 11. Displacement of Scratch Due to Creep Shown as (a) Raw Captured Image in Negative Format. (b) Enhanced Image for use in Datathief. 53 3.3 DATA EXTRACTION TECHNIQUE 3.3.] Image Enhancement After an image has been captured, it was edited in Adobe Photoshop®. Several tasks were performed on each photo before they were ready for analysis. Each digitally captured image was first inverted to a negative format. A white image with a black scratch was more convenient for analysis in Datathief than a black image with a white scratch. As the creep test progressed, the scratch became more difficult to detect asillustrated in Figure 11. A scratch which was only slightly visible in Photoshop was almost impossible to distinguish when opened in Datathief It then becomes necessary to highlight or trace the scratch with the line tool in Photosh0p®. This is illustrated in Figure 11. Most of the error involved in these creep tests probably occurs during this procedure, since the resolution of the scratch decreases with increasing displacement. In Figure 11, images in the left column are examples of digitally captured scratch displacement images (negative format) from a composite solder creep tested at 125°C. Time of capture is noted next to each image. The images in the right column have been enhanced in Dat'athz'ef as described previously. The images used in Datathz'ef must be of uniform dimensions. This is accomplished by establishing a uniform canvas size for all enhanced images. This eliminates the possibility of non-uniform image distortion in Datathief Both solder/substrate interfaces can be seen in the enhanced images and are highlighted in the image captured after 20 minutes. If the full raw image is opened in Datathz'ef, the area of interest is too small for accurate analysis. For these reasons each image is carefully cropped. However, as the 54 creep experiment progresses, the area of interest from each image increases vertically. By changing the canvas size or image border, the dimensions of the images stay the same while the surrounding ‘canvas’ is enlarged. Since subsequent images only grow in the vertical direction, the simplest process is to edit the final, tallest image from the creep experiment first. Sufficient borders are required on every image for the ease of data extraction. By examining the final image, appropriate dimensions can be determined and applied to all the previous images, with the first image having the largest vertical border. 3. 3.2 Data Extraction Datathief is a data digitizing shareware package available only for the Macintosh platform, which allows the user to extract data points from a digitized image by establishing an x-y coordinate system directly onto the image. Since the displacement of the scratch must be measured by extracting data from many images, it was necessary to establish the same x-y coordinate positions on each subsequent image. Since shear displacements were measured, a normalized dimensional scale based on the thickness of the joint was used as a basis for computing shear strains (Figure 12). The displacement of the scratch was modified or scaled with respect to the thickness in order to obtain a value of zero for the maximum displacement. See Appendix B for a detailed example of the data extraction process. The axes must be established in a manner so they can easily be accurately transferred to each subsequent image. The technique developed in this instance, was to use an 8.5 “ by 11” transparency overlaid onto the computer monitor. Several images from the same creep test, from beginning to the end, were then opened in the Datathief program and analyzed for similarities in an attempt to find an identifying 55 II" Profile of a shear-lap solder joint fabricated from Cu dog—bone halves I] Cu substrate ‘SnrAgm Cu substrate 6 0:5 "1 Normalized position across solder joint. Figure 12. Normalized Position Across the Thickness of a Solder Joint. ' 3'; me can options - ' ‘ The x and y axes were established using a pre-drawn coordinate system on a transparency which was overlaid onto the computer monitor. This allowed for the same coordinates to gm. be established for :5 1i each enhanced image. iii. Consistent units must we “Mm “memmm Image from Figure 11b image. (42 minutes elapsed time) opened in Datathief for data extraction. x = 0 ”at Figure 13. Screen Capture of Data Extraction Using Datathief 56 mark visible on each image near the stationary interface which could be used as a landmark throughout the data extraction procedure. This could be anything from a scratch to a small pore, but it must be visible on each subsequent image. A marker pen was then used on the overlaid transparency to trace this landmark as well as the left and right interfaces. These highlights can then be confirmed by opening different enhanced images, as seen in Figure 11b, in Datathief and checking for alignment of the transparency and the image. Once this was carried out, axes could then be drawn onto the transparency with the aid of a ruler. The longer the axes, the less error was involved due to the angle of View and parallax distortion inherent in the computer monitor. Figure 13 is a screen capture of an enhanced image seen, in Figure 11b, opened in Datathief. Once the axes are established, an x-y coordinate window in Datathief displays the current position of the cursor. The cursor was traced along the length of the enhanced scratch via the mouse. The mouse was then clicked at even increments, usually every centimeter or two on the monitor according to the pop up x-y coordinate window which indicates position on the image (Figure 13). The exact x—coordinate positions were noted and used on each subsequent image. The process was then repeated for the remaining images. Since it was impossible to determine the duration of each creep test beforehand, images were captured frequently, although well defined creep curves could be produced without extracting data from every image. Once the coordinate axes were established, about 10-20 datum points were extracted from the scratch, at the same x-coordinate positions for every image. 57 3. 3.3 Reproducibility of Data Extraction In order to determine how sensitive the data were to human interpretation, comparisons were made of a set of data extracted from the same set of images several months apart using the same procedure. The error between the two measurements of the same average strain increased with strain from about 1 to 10%, due to the degraded clarity of the captured image at larger strains. 58 3.4 DATA PROCESSING From the extracted data, a single plot can be produced which illustrates the entire scratch displacement history on one specimen (Figure 14). This is similar to overlaying all of the captured images onto one image, but in the form of a plot. Global creep strain is defined as the average creep strain across the thickness of a specimen. Measured global strains generally result in creep curves of the classic sense as shown in Figure 1. This designation is made since a comparison between the two forms (global and localized) is I'ICCCSSEII'y. 3. 4.1 Global Creep Strain Measurement The first step in calculating the global creep strain was to measure the shear displacement of one side of the specimen with respect to the other. Displacement during creep deformation occurred only at one solder/substrate interface, while the other interface remained stationary. A coordinate system was established by designating the end of the scratch at the stationary solder/substrate interface as a reference origin. The next step was to calculate the net shear displacement. In order to account for the non-horizontal shape of the initial scratch, the difference between the data for the initial scratch position and a simplified horizontal origin was subtracted from the subsequent scratch position data derived from images captured during the creep test. The resulting global net shear displacement after 8520 seconds was plotted with position across the thickness in Figure 15, where the bolder lines represent the net shear displacement. From this figure, the global creep strain was computed by dividing the total net displacement by the thickness. 59 Non-Composite Sn-Ag at Room Temperature, 21.4 MPa Scratch Displacement Due to Creep (Sample 039) ‘r I l l i i T T I l l I I I i 1.2_rrur Normalized Position . ' g E 5 . .' - i ‘ i : ' I ‘ , _ r~ . ; , . / ........... .._; 0.4 ,' - 1 J ; , . g : Before Creep - - -"..;=- - 4080 see ,I’ /' g 3 —B— 240 sec - O— 5640 see E] ___________________ g ________________________ 3 ____________ -l— 600 sec --z’ 7140 see , 0 2 1’. x /' 3 —V— 1260 see - -ir— 8520 sec E, E r --V--3120 sec 0 - . . . l r . r . . . . l r i . l . r i O 0.2 0.4 0.6 0.8 1 Normalized Position Through Solder Joint Thickness Figure 14. Scratch Displacement History. Global Net Shear Displacement (Sample 039) 1 .2 l j W ! I I I I I l T I I l ‘ f l I J b ' i - ; 4; 1 :_ ................................................................... i .......................................... . ............. J val-E .............. - .9‘ f _______ . ............. + --------- 9 ' " g ”filo-J" : 2% BE ------ . .....:::'.:’ _ $08 T """"""""""" """" [,LQQ‘i'Z""'i """"""""""" j o e ‘ ‘ §_.-- ' ’ I S — : i ,.--";" g _ 3 0.6 ~ ------- , - --------- ~ ------------- ; -------- _ .. ------------------------ 5 rrrrrrrrrrrrrrrrrrrrrr a D ’— ' ...o". , ’ I i - U .-.n T I I f . a o -. . ,.- , , , . . - .5 3 .e-tf.-_____ x _________ . . l a 0-4 j“ ' " ' ' ' “ ' ' ' " ' ’ ";”_j_;.;é" , , ’ g Before Creep (Simplified) - E ’ , ’ ’ 3 0 Before Creep (Localized) - g — _____ 3 , , , ’ § m-E- Before Creep (Global) ~ 0.2 3‘ - a" """""""" ’“y‘f' """"""""""""" 3' """"""" —-V-— 8520 see (Localized) i , , ’ 3 § - G - 8520 sec (Global) 3 _ , , ' ’ i 3 --------- 8520 see Net Shear Displacement 0 l r r l i 1 l L l l f x I r r x r 1 1 1 0 0.2 0.4 0.6 0.8 1 Normalized Position Across Solder Joint Thickness Figure 15. Difference Between Global and Localized Deformation. 60 3. 4.2 Localized Creep Strain Measurement The local shear strain was determined by fitting a cubic polynomial function using Kaleidagraph® software to the shape of each scratch. The same procedure for determining the net shear displacement was employed. This resulted in the localized net shear displacement shown in Figure 16. By differentiating the resulting cubic polynomials, the local SIOpe of each scratch, which is the local shear strain, was computed. Note that in some locations reverse (negative) strains were observed, such as position 0.7. The degree of accuracy of each fitted curve was indicated by the coefficient of determination, R2, where a value of 1 indicates a perfect fit. The average value of R2 for all of the curve fitting procedures in this experiment was 0.99. A plot indicating the creep strain history for different positions across the solder joint thickness (indicated by normalized position) was then produced from this information as illustrated in Figure 16 (strain-time history at selected illustrative positions is shown, note that at position 0.7, the strain rate is negative at later times). 61 1.2 Normalized Displacement The slope can be computed Simplified Origin 0.2 i __________ for any posmon 'x' once the ______ + Before Creep (Fitted) net displacement has been ——B—- 8520 SEC (Fitted) r.” COmPutéd --------- 8520 see Net Shear Displacement 0 l 1 L i 1 1 1 1 1 L l I 1 1 1 I l I l 0 0.2 0.4 0.6 0.8 1 Normalized Position Across Solder Joint Thickness Figure 16. Cubic Polynomial Curve Fit. 25°C, 21.4 MPa, Non-Composite Sn-Ag, Polynomial Fit 4 Localized Strain vs. Time (Sample 039) 1 I 1 f T I I I T I I I I r T E + 0 0.6 Normalized ' § j r 1} 0.1 A 3‘." Position : q 0.2 08- Through - 3 ‘ v: 0.3 />- ~ 0.9 Solder Joint """""""""""""""""""""""""""" g """""" j 04 to Thickness Er E (>- 05 Global 5 , / ’g ~ . , a ; 2 __ ................................................................................. ’./ ............................... _. l i _ t: r . ’ r E * i ‘ 1 _ . V y _ 0 _ . ...................... ‘ — ——————————— CT I A A A _ _1 l r l i l 1 l l l l 1 l 1 1 l l 7 0 2000 4000 6000 8000 Figure 17. Localized Net Shear Displacement (Sample 039) I I j l I I F l I I I I I I I I I I .. . . : j : I ‘ _ 5 2| W ‘“ _ ................................................ nee-.22---.-..--.---.--ff'.ft: ...... --.r ______ “ . . a 4; — - u ......................................................... Time (sec) Localized and Global Creep Strain. 62 3. 4.3 Strain-Rate, Position, and Time In order to examine how the strain-rate varied with time and normalized position, three- dimensional plots were produced fi'om the localized strain history plots. Polynomial curves of order 3-7 were used to fit the strain time curves (Figure 17) at each position. The derivative was calculated with respect to position to obtain the strain-rate, and plotted using Deltagraph® 3.0 software as illustrated in Figure 18. The 3-D surface is represented as a contour plot on the top of the plotting space. The shape of the 3-D surface varied from specimen to specimen. The example shown in Figure 18 illustrates the variability in the detailed strain-rate history in the measured part of the specimen. 63 3x10 Strain ZXIO Rate (1/5) 1x10’4’ —4 / 4/ z / Oz 1 Normalized Figure 18a. Figure 18b. Position 14.42 8.42 2 42 Time 0 ' (Klloseconds) 3x104 2110‘1 1x104 0.00015 0.0001 0.00005 0 Variation of Strain-Rate to Time and Normalized Position for a Composite Solder Creep Tested at 25°C. Normalized 0-5 Position (Seconds) 0.0004 0.0003 0.0002 0.0001 -0.0001 -0.0002 Variation of Strain-Rate to Time and Normalized Position for a Non- Composite Solder Creep Tested at 85°C. 64 3. 5 SEM OBSER VA T I ONS 3. 5.1 Determination of Stress After the specimen failed, the fractured halves were examined optically or with a SEM, and low magnification images were taken of the entire fracture surface, as shown in Appendix C. All specimens exhibited porosity, and the area of pores larger than 500 ,um2 were subtracted from the entire surface area, using NIH image software, in order to determine the actual cross-sectional area of the soldered region. The true stress operating on the cross sectional area was computed using this reduced area. Due to the long moment arm effect of the two copper tabs, strain in the joint was nominally uniform in the straining direction even though porosity present throughout the solder joint could lead to asymmetry in load carrying area with respect to the center-line. 3. 5.2 Initiation of Fracture The method employed for creep testing the miniature solder joints did not allow for detection of crack initiation. In order to determine the location of fracture initiation four solder joints, two fabricated from composite solder and two from non-composite solder were subjected to high stress dead load conditions. The solder joints were prepared in the same manner as described in Section 3.1.3. Visual records of the polished surfaces were recorded prior to loading with the use of a SEM. The solder joints were strained using the miniature creep frame. Each solder joint was initially subjected to a weight low enough to avoid rupture, approximately 2.27 Kg. (5.0 1b.). The load remained on the solder joint for roughly 10 seconds and was then removed. The solder joint was then removed from the creep frame and examined under an optical microscope. If no signs of 65 crack initiation were visible, the process was repeated with load increments of 0.23 Kg. (0.5 1b.). This procedure resulted in differing strain-rates depending upon each specimen. The strain-rates were not measured. When significant changes in the polished solder joint were detected, examination was performed with an SEM. 3. 5.3 Location of Fracture Examination of the fractured solder joints was performed to determine the location of the fracture across the polished thickness of the solder joint. These images were compared with the fracture surfaces images in order to better characterize the mode of fracture. 66 CHAPTER 4 RESULTS 4.1 COMPARISON OF EXPERIMENTAL CREEP DA TA T 0 DA TA FOUND IN THE LITERATURE In order to determine the reliability of the testing methods used in this investigation, a comparison of the creep data acquired was compared to data found in the literature. The data used for the comparison came from an investigation by Darveux et al. [41]. They subjected aged, realistically sized solder joints (ceramic chip carriers soldered together with eutectic Sn-Ag) to creep deformation at temperatures of 131°C, 77°C, and 27°C. Solder joints used in this study were not subjected to aging processes. The data of Darveaux et al. was normalized with respect to temperature in a manner described in Section 2.4.9. The normalized secondary (or minimum) creep strain-rate was plotted against normalized shear stress. Darveaux et al. used a ‘trial and error method’ to produce a best fit curve to represent the normalized creep data. Values determined by Darveaux et al. for G = G, + G1T(°C) , where Go= 2.8x106 psi and G1 = 1.0x104 psi/°C were assumed for this investigation. In order to simplify the comparison, a single plot comparing the data from this investigation to the data of Darveau et al. was produced (Figure 19). Another reason for this comparison was to examine the effects that the intentionally added Cu68n5 particles had upon creep behavior at room and elevated temperatures. From this comparison, it is apparent that the composite solder is more creep resistant than the 67 Normalized Creep Strain vs. Normalized Shear Stress . ' ' l g 103 E —— Darveaux ........................................................................ i E A 25°C Composite ................................................................... a 101 E ‘ 25°C Non-Composite ....................................................................... a. E 85°C Composite .................................................................... 101 i I 85°C Non—Composite ------------------------------------------------------------------ E ._\ O. or 125°C Composite 125°C Non—Composite y(TlG)exp(QlkT) _\ 0. (It Ill] 1 _X 0. \I Figure 19. ................................................................................................................. 6104 810'4 10'3 rIG Comparison of Secondary Creep-Rate Data. 68 Normalized Creep Strain vs. Normalized Shear Stress 103 E Darveaux I __________________________________________ 2 E A 25°C Composite ....................................................................... g A 10‘ E ‘ 25°C Non-Composite ------------------------------------------------------------------------ E E E 85°C Composite """"""""""""""""""""""""""""""""""""""""""""""""""" “E g 10'1 E ' 85°C Non-Composite """""""""""""""""""""""""""""""""""" j: g ‘3 E 125°C Composite """""""" """" v """""" A """""""""" 1 610 E 1250C Non-Composite EEEEEEEEEEEEE v EEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEEE - F A .......................... 3:;- 10'5 e ----------------------------------------------------------------------------------------------- a 210‘1 4104 6104 810“1 103 13,6 Figure 19. Comparison of Secondary Creep-Rate Data. 68 non-composite solder, at room temperature and 85°C, but the composite solder joints are comparable in creep resistance to the non-composite at 125°C. 4.1.1 Room Temperature Composite Data The composite solder joints tested at room temperature exhibited superior resistance to creep as compared to the non-composite solder joints. The strain-rate of the solder joint tested at the lowest stress was near 3.5 orders of magnitude less than the trend established by Darveaux et al. As stress increased this difference decreased by an order of magnitude on average. A trend different from that established by Darveaux et al. for the non- composite data appears to be present for the composite data resulting from this experiment. 4.1.2 Room Temperature Non-Composite Data Room temperature (25°C) creep data obtained in this study was in relative agreement with the trend established by Darveaux, et al. The three solder joints tested at higher stresses were all characterized by strain-rates within an order of magnitude of Darveaux’s best fit curve. Two of the points were in such proximity that only one was visible. Of the two solder joints tested at lower stresses, both were characterized by strain-rates which were lower than the established trend. One solder joint had a strain-rate within an order of magnitude of the established trend, while the other specimen had a strain-rate more than an order of magnitude less than the established trend. 69 4.1.3 85‘C Data The two composite solder joint tested at lower stresses agreed well with the best fit curve. Both were characterized by strain-rates well within one order of magnitude of the established trend. As stress increased, the margin between the data and the best fit curve increased to around one order of magnitude. Non-composite solder joints tested under this temperature environment exhibited similar results to the composite data. 4.1.4 125‘C Data All of the composite solder data was characterized by strain-rates within one order of magnitude of the trend established by Darveaux et al. Two of the solder joints tested exhibited higher strain-rates than the established trend. The other two specimens were characterized by strain-rates which were equal or almost equal to the established trend. Both of the non-composite solder joints tested at this temperature had strain-rates nearly equal to the established trend and their data points fell almost directly upon the best fit curve. These samples appeared to have slightly better creep resistance than non- composite solder joints tested under lower temperature conditions. However, this apparent trend should be disregarded due to the inherent scattering of data associated with creep. 7O 4.2 GLOBAL VS. LOCALIZED CREEP STRAIN The global strain-time history usually exhibited classical primary, secondary and tertiary strain behavior as indicated by the heavy lines in Figure 20b. The vertical scale of a global creep curve became flattened when plotted with the localized creep curves from a single specimen since a wider range of values were present. As a result, plots of individual global creep curves are available in Appendix D. Characterization of the local strain history of all samples included in this investigation is provided in the following sections. As mentioned previously, these specimens include composite and non- composite solder joints deformed at room temperature, 85°C, and 125°C. Since there are more than 20 solder joints tested, descriptions will be limited to the most representative specimens. The associated figures include a scratch displacement history (a), a strain history (b) and strain-rate history with respect to position (c). 4.2.1 Room Temperature Composite Creep Data Figure 20 illustrates creep history of a composite solder joint subjected to creep deformation under a stress of 26.2 MPa for about 4.5 hours (16.1 Ks). This specimen had highest localized strain accumulation near the interface at position 1.0. Identification of the beginning of global tertiary creep was near a shear strain of 0.5. However on a local scale, tertiary creep appeared to start at a larger shear strain of 1.2 near the interface. The secondary creep-rate was measured to be 4.2x10'5 5". The 3-D plot indicated a sharp initial decrease in the strain-rate near the interface (positions 0 and 1.0), with a much lower initial strain-rate in the center of the specimen. 71 25°C, 26.2 MPa (Sample 053) Composite Sn-Ag 25°C, 26.2 MPa (Sample 053) Composite Sn-Ag Localized Strain vs. Time 2.5 ' _ r 1 i I 02 vNormalized 3 .‘ 04 Posmon ,7 E 0-8 Elapsed 2 ,1 06 Through ..9 o ‘ - ’ ' Solder Joint ,/1 E Tlme y 08 _ I, I a) . . (Sec) ,_ Thickness . L; g 2 —D—0 1.5 -,.,--- 1.0 ------------- w ----- u, 5 /n" _ ,‘O —B— 420 .5 Global "504 "";i"“ """ r " 2340 g . ; 'o ,..(; o , v 5580 a; 3 La ....... 7500 '7; 0'2 ” .0 + 9660 E . 11580 g o -------------------------- . . <> . 16140 I -I -i i i _o'2 llilllil lll]|| 0 - llllllllllll v v C O O O *0 O O O O 0.2 0.4 0.8 1 o 8 8 8 8 ._ c: z 8 Normalized Position Through N " ‘° °° ‘— x— .4 .— Solder Joint Thickness Time (sec) (at) (b) / fie 0.00035 // \E/W 0 0003 / \ ' 0.30 , A ><\ \’ 0.30 . \ ’ ' \ \\, 0.00025 \ \\ Strain 0.201 \\ \r0v20 0.0002 Rate \\ (US) l \\ \' 0 00015 \ l ‘ 0.10 ’ 7C0 10 , 0.0001 0 E , 0 0.00005 1 14.42 Normalized 0 Position 8.42 Time 0 2'42 (Kiloseconds) (C) Figure 20. Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 25°C. ra pC Fi de 5P ho fof CR 512 be be P0 AI Val Cnc The strain-rate decreased with time in a normal way to reach a minimum strain-rate in the range of 2-8 x 10'5 s". The strain-rate then gradually increased to reach higher values at the two interfaces prior to rupture. The highest strain-rate (3.3 x 10'5 s") was observed at position 1.0. 4.2.2 Room Temperature Non-Composite Creep Data Figure 21 illustrates creep history of a non-composite solder joint subjected to creep deformation under a stress of 11 MPa for a duration near 2 months (6.0 Ms). This specimen had highest localized strain accumulation near the interface at position 0, however the specimen deformed in a homogeneous manner with the localized strains following closely to the trend of global strain. Identification of the beginning of tertiary creep is near a global shear strain of 0.15. Near the interface, tertiary creep appeared to start at a slightly higher shear strain near 0.18. The secondary creep-rate was measured to be 3.0x10'8 s“. The 3-D plot indicated a sharp initial decrease in the strain-rate near the interface (positions 0) and to a much lesser extent position 1. The strain-rate initially peaked between positions 0.3 and 0.5 peaks at a value near 5.0x10'8 s". The strain-rate at position 0 remained constant throughout the test after a slight increase near the beginning. At the other interface, position 1, the strain-rate rapidly increased where it peaked at a value of 7.5x10'8 5'1 afier 2 Msecs. It then decreased, followed by a final increase at the end of the test to a value of 2.0x10'7 5". From approximate positions in the range 0.8 to 0, the strain-rate remained relatively constant until the end of the test, where an increase 73 25°C, 11 MPa (Sample 042) Non-Composite Sn-Ag .0 A .0 w ......... .0 N Normalized Displacement O liillllllllllllllll O O 0.2 0.4 0.6 Solder Joint Thickness (a) 0. Normalize Position 0.8 1 Normalized Position Through 25°C, 11 MPa (Sample 042) Non-Composite Sn-Ag 0Spratch Displacement Due to Creep Localized Strain vs. Time . lllIlllllllllllE‘lTl 03 . r1]r1rr_ . , z._. 2320:?“ s a 0.6 ******************************** 0.25 “EH ‘0’: Through / Elapsed , 09 Solder Joint I / ; 0.5 ---------------------------- Time 02 1.0 Thickness , — (K390) Global 5 ”1 ............................ llllllllll .\;\c o r - _005 Plll|llllllllllllllllllillllli_ (D (D (D ‘0 (D ‘D O O O O O O O 1- ‘— ‘— v— ‘— x— F N ('0 V 10 L0 Time (sec) 0)) 0.0000002 0.00000015 0.0000001 ;. 0.00000005 000000005 00000001 000000015 0.0000002 000000025 00000003 (Kiloseconds) (C) Figure 21. Displacement (a), Creep Strain (b), and Strain-Rate (c) Histories for Non- Composite Solder at 25°C. 74 [E \\ re 51: CC th F1", de SD in strain was experienced across the entire solder joint with the maximum strain—rate at position 1. 4.2.3 85 °C Composite Creep Data Figure 22 illustrates creep history of a composite solder joint subjected to creep deformation under a stress of 13.2 MPa for about 1.3 hours (4.6 Ks). This specimen had highest localized strain accumulation near the interface at position 1.0. Identification of the beginning of tertiary creep was near a global shear strain of 0.5 but near the interface, tertiary creep appeared to start at a larger shear strain of 0.75. The secondary creep-rate was measured to be 9.8x10'5 3". The 3-D plot indicated a slight initial decrease in the strain-rate near the interface at position 0, with a gradual increase in strain-rate towards position 1. The strain-rate remained relatively constant along position 0 except for a slight increase and decrease in strain-rate at the start and finish of the test respectively. At position 1, the strain-rate continually increased to peak at a value near 8x10'4 3". The strain-rate remained constant throughout the duration of the test at position 0.2 and increased towards position 1 throughout the test. 4. 2.4 85 °C Non-Composite Creep Data Figure 23 illustrates creep history of a non-composite solder joint subjected to creep deformation under a stress of 19.6 MPa for about 34 minutes (2.0 Ks). Unlike the prior specimen, the largest strains were observed in the center of the specimen near position 0.5. Identification of the beginning of tertiary creep is near a global shear strain of 0.24 75 2% Fi 85°C, 13.2 MPa (Sample 033) 85°C, 13.2 MPa (Sample 033b) Composite Sn-Ag Composite Sn-Ag Localized Strain vs. Time, ngratch Displacement Due to Creep 1 5 ' : : : I ; ' + 0 Normalized ' I i L i : Position , 4 «0'7 : l ' y H7; 8; Through / . «l 506 g .[ i _.x’_’,-,:@; 1 v 00 S°'der A , % GE, ' ; . ' ' 31 0.4 . g 4 . v .. 2P 0.6 , ; 1 163-05 (E 7": . ,; o 0.8 ~ 3 ~ 1 5 ..5; ~-— 1.0 .. 3:7... 5,-0.4 1— V V ' =0.5 Global ' _. c" . 1 1: ’__ - - , 1 i 5 A ‘ “0.2 ' o i . i -~1 E :L ‘ .0 E : ' 1 : :1 20.1 t , ----- ~ ‘ 5 3 : i ' 0’ 1 . a I ; . i 4 0 i Yer-1' L i‘l‘l‘ ‘ I ‘ _05 WW 0 0.2 0.4 0.6 0.8 1 0 § § § § § Normalized Position Through T N " “’ Solder Joint Thickness Time (see) (a) (b) 6x10“ 6x104 4x104 1 4x104 Strain , Rate 4 4 (us) 2x10 1 2x10 0 ’ 0 -2x104 , -2xl0“ 1 4820 Normalized 0.5 3120 Position 1820 Time 0 (Seconds) (C) Figure 22. Displacement (a), Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 85°C. 76 85°C, 19.6 MPa (Sample 025) 85°C, 19.6 MPa (Sample 025 Non-Composite Sn-Ag Non-Composite Sn-Ag Scratch Displacement Due to Creep Localized Strain vs. Time 0.35 TIVITTTIIlTTITTllTI 0.5 1W7: : ‘ 3 ——a—— 05 v o 9 f}? E .. 03 ________________ _- 0.4 —a— 0.7 ~<> 1.0 w ------ —_ c ' I 0 8 Global . ‘ i” 3 0 3 _ ......................... . ............. _‘ §025 ' g ; :F 5 (I j 2 :_ ................. . . . . . . ..... L E. 0 2 S Elapsed 0 : : {/21 3 . _ _ c , a g - Time '3 0 1 _ ............. . ..... .; 1 (Sec) h ' 3 , v i 0.15 a : ‘ ' ' V v . _ E -e~ 1080 a) 0 ”$14): Y...‘ r; v.1 ' 'V i = 01 _ —e— 120 --|— 1320 3 ; ”x j I N - t 9' 480 . 1680 -01 .— """ ’" " -~>\-..\~;» i E H _ v 660 e 1860 g ; . a : z°0.05 ------ +760—0—2040 -o_2 5 ------- fie —: ' —t—- 900 Z : 5 1 ‘ : 0 111 '0.3 ’llllll lI_L_LAlillllel 1‘ o o o o o 0 0.2. 0.4 0,-6- 0.8 1 o 8 g g g g Normallzed Posmon Through “ “ N N Solder Joint Thickness Time (see) (a) (b) =§§é ' ain‘t 6x104 1 gal! 6x104 - Elli .1- “>39. 4x104 4X10 -i=‘ Ill? . 2x104 ‘ "' 2x10 Strain .. : Rate 0 . 0 (Us) -2x10'4 ‘ -2x10“ -4xl0““ 4x104 Normalized Position 0 520 Time (Seconds) (6) Figure 23. Displacement (a), Creep Strain (b), and Strain-Rate (c) Histories for Non- Composite Solder at 85°C. 77 Fig def ach ldel Spe. the Cl'Cc' Posi but near the center, tertiary creep appears to start at a slightly larger shear strain of 0.3. The secondary creep-rate was measured to be 1.2x10'4 3". The 3-D plot indicated a slight initial decrease in the strain-rate near the interface at positions 1 and a higher initial strain-rate in the center of the specimen. Unlike the prior specimen, the strain-rate remained the highest at the center of the joint. At position 1 the strain decreased until specimen fracture. Then strain-rate near position 0 increases until an elapsed time of 1400 seconds where it attained a maximum strain near 2.5 x 10‘1 s'1 and began a decrease in strain-rate until the end of the test. The highest strain-rate (5.0 x 10‘1 3") developed in the center of the specimen between positions 0.3 and 0.5 at the end of the test. 4. 2.5 125 °C Composite Creep Data Figure 24 illustrates creep history of a composite solder joint subjected to creep deformation under a stress of 11.6 MPa for about 18 minutes (1.1 Ks). This specimen achieved highest localized strain accumulation near the interface at position 1.0. Identification of the beginning of tertiary creep for this specimen was not obvious. The specimen behaved in a tertiary manner after a global shear strain value near 0.1, but near the interface, tertiary creep appeared to start at a larger shear strain of 0.3. The secondary creep-rate was measured to be 5.8x10'4 5". The 3-D plot indicated a sharp initial increase in the strain-rate near the interfaces positions 0 and 1.0, with a much greater increase near position 1 (5x10'3 3"). A sharp decrease in initial strain-rate initially existed at the center of the specimen with a value of -2x10'3 5". Near position 0, the strain-rate remained relatively constant until a decrease 78 “cartoon-00.0 UeNszscz 125°C. 11.6 NIP-at (Sample 028) 125°c, 11.6 MPa (Sample 028) CompOSIte Sn-Ag Composite Sn-Ag (Spratch Displacement Due to Creep Localized Strain vs. Time .ir 1 . . .. 2...”; ; . l‘ 5 0'6 [E W E ' : ' ' 1 Elapsed 0.8 --: -* , i g - 1' " ‘ Time i 005 (See) ,’ , , . g - —0—0 3 3'04 >B— 120 .E x Q 480 u 0.3 720 g 0'4 fl 3 . 840 i = + 900 E 0'2 960 ° 2 4 I— 1020 ‘2’ 0.1 1080 0 , . : _i 1140 1 : - . . 0@;_o 11_l lli‘lllllllllllllJ_1_LJll 0 0.2 0.4 0.6 0.8 1 0 § § § § § § Normalized Position Through _ “ " Solder Joint Thickness Tlme (see) (a) (b) 7. <\ \. ‘ Z>r 0004* : \\w0.004 .i/Hpv \ \\r 0.0025 \\— 0.002 Strain 7 \- Rate ’ (Us) 0 —o a \\.L 0002' \\~-0.002 ' 1020 Normalized 0.5 620 Position 320 Time 0 (Seconds) (C) Figure 24. Displacement (a), Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 85°C. 79 If before p r0 ryh A6 €hy cyl eyr au€hy pa’cyr jacl e, ehy pirdcl -rdey C1 crydpyC ‘hr ra’ jh’y d third of the test duration, remained constant for another third, and then increased again to terminate in a manner completely opposite of the initial strain-rate. The entire solder joint behaved in an opposite manner as it did at the beginning of the creep test achieving a 1 maximum strain-rate of 1.5x10'3 5’ . The strain-rate near position 1, sharply decreased and then behaves erratically until a continual decrease half way through the test. 4. 2. 6 125 °C Non-Composite Creep Data Figure 25 illustrates creep history of a composite solder joint subjected to creep deformation under a stress of 12.7 MPa for about 50 minutes (3.1 Ks). This specimen achieved highest localized strain accumulation near the interface at position 0. Identification of the beginning of tertiary creep was near a global shear strain of 0.6, but near the interface, tertiary creep appeared to start at a larger shear strain of 0.9. The secondary creep-rate was measured to be 1.3x104 s'l. The 3-D plot indicated a relatively constant strain-rate across the thickness of the solder joint with a slight decrease at position 0. The strain-rate remained constant throughout most of the test except at position 0 where the strain-rate appears to follow a sinusoidal pattern until rupture. After an elapsed time near 2600 seconds, the strain-rate sharply increased in a uniform manner, except near position 1, with the maximum value of 2.2x10-3 s-l at the center of the solder joint. The strain-rate near position 1 remained relatively unchanged throughout the duration of the test. 80 125°C, 12.7 MPa (Sample 027) Non-Composite Sn-Ag Scratch Displacement Due to Creep .-E1—"’—' .. i 1.2 50.8 Elapsed E ' I Time 1 8 1 1 (Sec.) (5 ’ j 30.6 . V i ...... '2 1 'li , o .. . ,1‘ 30.4 - g , ...... 5 ‘ ,' : Tu ' E02 . 5v ......... O ,7 . . 2 I, : 125°C, 26.2 MPa (Sample 027) Non-Composite Sn-Ag Localized Strain vs. Time —D—0 -—*——0.7 —B— 0.1 .,... 0.8 0.2 ...i 0.9 10 Global '1 0 0.20.40.60.81 o § § § § § § Normalized Position Through " " N N "’ Solder Joint Thickness Time (sec) (a) (b) // 0.0025 - —0.0025 0 0025 >< ’ 0.002 // 0.0015“ . 0.0015 // 0.0015 Strain Rate / 0.001 (l/s) - , 0.0005 0 . _ 0 0.0005 -0 0005 — - 00005 0 3080 Normalized Position Time -0.0005 0 (Seconds) (C) Figure 25. Displacement (a), Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 85°C. 81 \‘2 (11 ii 4.3 POSITION 0F MAXIMUM LOCALIZED CREEP STRAIN It is evident through determining the localized creep strain, that strains of much higher magnitudes than the global strain are present at different positions across the thickness of the solder joints. The room temperature non-composite solder joints creep tested at higher stresses all exhibited maximum localized creep strains along the interface. The non-composite solder deformed at lower stresses yielded higher localized strains near the middle of specimen. Non-composite solder tested at 85°C resulted in two samples exhibiting maximum localized creep strain at the interface, and the other sample very near the interface. Composite specimens tested at 85°C yielded mixed results with half achieving maximum localized strains at the interfaces, and half near the middle of the solder joint. Non-composite solder tested at 125°C all exhibited maximum localized creep strain values along the interface. The composite specimens again yielded mixed results. Two of the samples behaved in an extremely homogeneous manner with the localized creep values very similar to the global creep values. One of the two remaining samples demonstrated maximum localized strain values at the interface, while the other sample yielded maximum localized creep strain values near the middle of the specimen. 82 4.4 Mic: 0011' 0001 W35 r00 sur ho: ele so SO 4.4 FRACTURE SURFACES Micrographs of the fracture surfaces were taken using an SEM. An attempt to correlate outlying positions on the Dom plot by analyzing the fracture surface was made but conclusive evidence was not found. However, a temperature/composition related trend was apparent. Fracture surfaces (Figure 26a) from non-composite solder joints creep tested at room temperature and low stresses, appear to be relatively smooth and flat (planar) with one or two big steps. With increasing stress, the fracture topography took on a rougher appearance with many levels of fracture surface elevation (Figure 26b). The room temperature composite samples (Figure 260) at higher stresses possessed smooth surfaces which appeared to be stippled rather than smeared, suggesting that a more homogeneous failure process occurred. Fractures surfaces of all specimens tested at elevated temperatures (Figure 26d), were similar to the room temperature non-composite solder at lower stresses. The fi'acture occurs mostly on one plane, but the surface was somewhat rougher, and more pronounced smearing was present than the non-composite specimens crept at 25°C. 83 Fl Fig 21.7 MPa. (Sample 038) 25°C, I = Non-Composite Fracture Surface, T Figure 26a. of.” a . M. 23.5 MPa. (Sample 040). 25°C, 1 = Composite Fracture Surface, T = Non- Figure 26b. 84 Fi Fig Figure 26c. Composite Fracture Surface, T = 25°C, 1: = 32.2 MPa. (Sample 052). Figure 26d. Composite Fracture Surface, T = 125°C, 1 = 9.1 MPa. (Sample 036). 85 4.5 RESULTS OFLOCATION 0F FRACTURE EXAMINATION The solder joints subjected to creep deformation in this investigation underwent fracture in a consistent manner. In general, two types of fracture occurred during this investigation, fracture through the middle of the specimen and fracture near one of the interfaces. Fracture appears to be dependent upon strain-rate and the amount of porosity present in the solder joint. If a large pore was present in the solder joint, approximately 20% of the fracture surface area, fracture occurred through the center or diagonally across the thickness from one interface to the other. Solder joints which lacked porosity fractured along or near one of the interfaces when subjected to high strain-rates (104 5"). However, when subjected to low strain-rates (10'8 s'l) fracture occurred through the middle of the solder joint for the specimens examined. No correlation between composite and non-composite solder was noticed, besides the fact that the composite solder generally possessed a greater amount of porosity. Results for the different testing conditions are indicated in Appendix E along with SEM micrographs of the fracture and fracture surfaces. 86 4.6 RESULTS OF CRACKINITIATION TESTS The results of this investigation were similar to the results in Section 4.5. In general, fracture initiated along or near the interface. Three of the four solder joints tested indicated an initiation at the interface. The fourth solder joint, fabricated from composite solder, demonstrated crack initiation near the center of the solder joint. Upon rupture of the specimen a large pore was present in the center of the joint, roughly 20% of the fracture surface area in size. It should be noted that all tests in this particular investigation were performed at high strain-rates, greater than 10'3 5". SEM micrographs for these tests are included in Appendix E. Small circular areas (Figures 27 and 28), varying in diameter from approximately 10 to 70 ,um were noticed to form near an interface after some deformation. These area were present on one solder joint prior to testing and are believed to be the results of strain during polishing. The areas appear to separate themselves from the rest of the strained matrix. Upon further deformation, the areas behave in two different manners. In some instances the areas tend to collapse into the matrix Figure 27c. The areas have also been observed to ‘pop’ out of the matrix as indicated in Figure 27b and Figure 28. Although these areas are a result of deformation, the exact cause is currently unknown to this investigator. Since they tend to form near the interfaces, they may be linked to the Cu68n5 intermetallics which naturally form there. 87 (C) Figure 27. Formation of Circular Areas Under High Stress Conditions. (a) Formation of Circular Areas. (b) ‘Pop’ Out Upon Further Deformation. (c) Collapse of Circular Area. 88 (b) Figure 28. Solder Joint Micrographs Before (a) and After (b) Further Deformation. 89 CHAPTER 5 DISCUSSION 5.1 CREEP BEHA VIOR A summary of all creep tests is provided in Appendix F. The creep behavior of the solder joints is best summarized in the Dorn plot in Figure 19, which shows the normalized shear strain-rate vs. the normalized shear stress. The solid line describes the non- composite Sn—Ag data of Darveaux et al. and it is apparent that most of the present data are in agreement with the trend established by this best fit curve. The addition of the Cu6Sn5 particles as matrix reinforcements significantly improved creep behavior at room temperature conditions. The room temperature composite data were as much as 1000 times more creep resistant than the non-composite solder. At 85°C, the composite data were in close agreement with the Darveaux data, but the non- composite creeped at least 10 times faster than the composite. At 125°C, the creep resistance of the non-composite and composite solders were similar. In addition to increased strength the composite solder appeared to deform in a more homogeneous manner 90 5.2 HOMOGENEI T Y 0F DEFORMA TION Determination of homogeneity of deformation can be examined using visual and analytical methods. Observations of the creep strain-time history plots conveyed a good sense of whether or not the sample deformed in a homogeneous manner. Local tertiary creep generally commenced only in a limited part of the specimens where the strains are highest. Analytically, by taking the ratio of the onset of global tertiary creep to the onset of localized tertiary creep, a value resulted representing the homogeneity of deformation. The closer the value was to l, the more homogeneously it deformed, since a ratio of 1 implies that the onset of local and global tertiary creep occurred at the same strain, which in turn implies that local and global strain-time histories are the same. Figure 29 shows how the onset of local tertiary creep varied with secondary strain- rate. Room temperature creep of the non-composite solder at higher stresses tolerated high local shear strains near a value of 3 before tertiary creep commenced, as compared to the composite solder, where the beginning of tertiary creep occurred at local shear strains in the range of 0.2-1.2. At lower stresses, the non-composite specimen exhibited local tertiary creep at a strain very close to the global tertiary creep strain. Since the composite specimens generally exhibited a local strain for the beginning of tertiary creep close to the global strain at the beginning of tertiary creep, this indicates that the strain evolved more homogeneously in the composite solder. It may also be an indication that strain evolves with lower stress in the non-composite solder, as indicated by the lines that suggest these trends on the various strain history plots. At elevated temperatures, localized tertiary creep began at strains between 0.2 — 1.5 for all specimens. 91 Figure 30 illustrates how the strain homogeneity ratio varies with strain-rate. A trend which suggests that homogeneity of deformation decreases with increasing creep-rate is indicated for the room temperature non-composite data. For room temperature and 125°C deformation, the composite solder exhibited higher levels of homogeneity than the non-composite solder. At 85°C, the trend is not as easily identified. The most homogeneous deformation occurred in the composite at 125°C, but the onset of local tertiary creep was at a lower strain than most other specimens as shown in Figure 30. The large amounts of strain tolerated by the non-composite samples under a high rate of deformation (lxlO'4 or greater) is indicative of the inaccuracies involved in accelerated testing. The toleration of such a large amount of strain before the onset of tertiary creep is unrealistic. A possible relationship between porosity and tertiary creep was also investigated. A plot was produced which compared homogeneity of deformation to the average percent porosity present in each fractured soldered half (Figure 31). The plot suggests no dependency upon porosity. Therefore, porosity was not a major variable on damage in this Study, 92 Onset of Local Tertiary Creep vs. Secondary Creep Rate 3 r, I I l I I I I I I I I l I I I I I f I I I I I I I ! T I I T r I ‘1 I I I i l T I Id 5 r A Composite 25°C 1 to C , .......................................................................... j (.3 2-5 L ‘ Non-Composnte 25°C . 3 : Composite 85°C j g 2 _ fl .......................................................................... z to 3 I Non-Composite 85"C . E; ; Composite 125°C 1 t L4 .. K 1' . ...... __ '93 1'5 r ° Non-Composite 125°C . j _ t . A ; ,. 8 L ' i a .._ , f g f o _ f # , , ,; -— ._ : £05 r #,##~0 """"" 3.; """""" V: O : A 3 A V : 0 i— l l l l l I l Ii 1 l l l l l J_1i l l l l I l I | l L 41 J._l 1 L4 Ii 1 l l J l I l l- 10‘8 10‘7 10's 10'5 0.0001 0.001 Global Secondary Creep Rate Figure 29. Onset of Local Tertiary Creep vs. Global Secondary Creep-Rate. 1 Homogeneity of Deformation vs. Secondary Creep Rate . .o .o (D (O IIIIIIIII .0 \l lIIIIIIIIl IIII] .0 01 l I I Onset of Tertiary Creep Ratio (Global I Local) O 0) I I I IITIIr I I IIIIII T I ITITIII v I l Composite 25°C Non-Composite 25°C Composite 85°C 0.4 r """""""""""""""""""" : V Non—Composite 85°C 0.3 E— 5.- Composite125°C T , 0 Non-Composite 125°C ‘3 j 0'2 '— I lillllli 1 LIIXIF 1 r LJIIIII l I lllllli l 1L41J_L1~ 10'8 10'7 10'6 10'5 10'4 10'3 Secondary Creep Rate Figure 30. Strain Homogeneity Ratio vs. Global Secondary Creep-Rate. 93 Homogeneity of Defamation vs. Average “lo Ffprosity I l 1 -\..ai T I 1 f T I T I fiI I I I I T‘T I I I T T F .2 i v 2 ‘5 ........... A ............ . ......................... Y ............................................................. _. tr 0-3 , 5 , r) , §= * : A ‘ cg z ........................................ L .................... .2 if 06 r 02 ; A 025°C - .5; : 0 V g 5 A NC25°C § _, Z ”'8 i—A. ....................... I ................... ‘ o .-..-.--..- ........... .4 T5 0.4 . 3 A ~ C850 ; Ev . 9‘ ‘ § V NC85°C ‘ 1 o L I ' O o _l g 02 __ .................. ' ....................... . .................... C125C L ....................... _ o . 3 g 0 NC125°C i 0 L 4 A; L I I L l I l J l i J I 1 I i l 1 l 1%] 0 0.05 0.1 0.15 0.2 0.25 Average % Porosity on Load Bearing Area Figme 31. Homogeneity of Deformation vs. Average Percent Porosity. 94 5.3 ANALYSIS OF FRAC T URE The solder joints subjected to higher strain-rates (10'4 3") all exhibited fracture along the interface, unless porosity was present in the amount of at least 20% of the fracture surface area. It has been observed that Cu6$n5 intermetallic particles naturally form at the Cu substrate-solder interface (Figure 6). The presence of these unavoidable intermetallics would have an embrittling effect on the microstructure as compared to the compliant solder matrix. As a result, the solder joint is weaker at the interface and will tend to fracture there unless significant porosity is present throughout the solder matrix or a much lower strain-rate is experienced. This manner of solder joint fracture was consistent with expected results mentioned in Section 2.3. 95 CHAPTER 6 CONCLUSIONS A criteria for damage based on the onset of tertiary creep was examined in shear creep deformation of miniature solder specimens made from eutectic Sn-Ag solder with and without 20v% 5-8 pm Cu68n5 reinforcements. The strain-time history was measured both globally and locally and the range of local strain variation was quantified. The composite solder exhibited a more uniform deformation, but tertiary creep commenced at a lower creep strain than non-composite solders at room temperature. The local shear strain where the onset of tertiary creep was observed increased with strain-rate for both solder types, and the fracture surface features became more rough. At elevated temperatures, tertiary creep commenced at a similar strain for composite and non-composite specimens, and the fracture surfaces were similar. The composite material was more creep resistant at lower temperatures, but it is not yet clear whether this benefit outweighs the fact that less strain is tolerated locally before tertiary creep commences. Porosity appeared to have no effect upon homogeneity of deformation. 96 APPENDICES 97 APPENDIX A CREEP MECHANISMS 98 Stress Regime n T p Q Mechanism 1 High 2 QL Naborro-Herring Low 1 Low 3 QGB Coble 1 High 0 Q, Harper-Dom 5 High 0 QL Dislocation Climb (M-Type) Intermediate 7 Low 0 Q j Dislocation Climb (M-Type) 3 High 0 Q1 Dislocation Viscous Glide (A-Type) 5 Low 0 Q i Dislocation Viscous Glide (A-Type) High * H/L O Q l or Q, Power Law Breakdown Adapted from Darveaux et al., page 539, and Yang et al., page 1137. Definitions Lattice Self Diffusion L QoB Grain Boundag Diffusion Qj Dislocation Pipe Diffusion 1 Inter-Diffusion of Solute Atoms Table l. Creep Mechanisms 99 APPENDIX B DATA REDUCTION 100 Table 2 represents extracted data from an enhanced scratch displacement image which was digitally captured via a traveling microscope. The data extraction was performed via the Datathief shareware program. The first column, ‘x position’, represents the thickness of the solder joint, from one interface to the other, as it appeared on the computer monitor under the Datathief program. Except for the first and last values, the data are extracted at equal intervals. The units are centimeters, i.e., the thickness of the solder joint as it appeared on the computer monitor is represented by a value of (15.78 cm - 5.11 cm =- 10.67 cm). The actual thickness of the solder joints ranged from 50 to 100 m. The remaining columns represent the position of the scratch line prior to loading of the solder joint and during the creep experiment. For example, the column labeled ‘2940 sec’ represents data extracted from the image captured 2940 seconds into the creep experiment. On each captured image, the data was extracted at the same x position established in the first column. It can be seen that as time progressed, the position of the scratch remained relatively constant for the first row of data, labeled ‘5.1 l ’, but changed drastically in the last row, labeled ’15.78’. 101 102 x position Before Greg) 1080 sec 1920 sec 2940 sec 4020 sec 5700 sec 5.11 8.81 8.77 8.64 8.64 8.72 8.61 5.98 8.87 8.80 8.33 8.01 7.77 7.86 6.99 8.98 8.71 7.81 7.14 6.81 6.56 7.99 9.12 8.57 7.26 6.37 5.69 4.94 9.00 9.26 8.37 6.72 5.53 4.59 3.12 10.04 9.40 8.13 6.11 4.66 3.40 0.84 11.04 9.55 7.90 5.56 3.94 2.27 -l.16 12.05 9.69 7.64 5.10 3.19 1.29 -2.72 12.99 9.83 7.38 4.64 2.41 0.22 -4.48 13.99 9.94 7.12 4.27 1.84 -O.50 -5.63 14.97 10.11 6.92 3.84 1.32 -1.17 -6.73 15.78 10.19 6.74 3.55 0.89 -1.51 -7.28 Table 2. Scratch Displacement Data. In order to simplify the data reduction process, the data representing the thickness of the solder joint was normalized. The thickness of the solder joint was normalized by subtracting the origin, 5.11, from each value in the first column and then dividing each of those values by the established solder joint thickness (15.78 - 5.11). This resulted in a normalized solder joint thickness ranging from 0 to 1. The scratch displacement data was modified or sealed according to the thickness of the solder joint. For simplicity, the maximum displacement of the scratch, represented here by a value of -7.28, was shifted to a value of 0 and then sealed to the solder joint thickness. This was done by adding a value of 7.28 to every scratch position value, i.e., every number in every column with exception of the first column. Each modified value was then divided by the solder joint thickness (15.78 - 5.11). The results are seen in Table 3. 103 104 x position Before Creep 1080 see 1920 see 2940 see 4020 see 5700 sec 0.000 1.507 1.504 1.492 1.492 1.499 1.489 0.082 1.512 1.506 1.462 1.432 1.409 1.418 0.176 1.523 1.498 1.413 1.351 1.320 1.296 0.271 1.536 1.484 1.362 1.278 1.214 1.145 0.365 1.550 1.465 1.311 1.200 1.112 0.974 0.462 1.563 1.444 1.254 1.119 1.000 0.760 0.556 1.576 1.422 1.203 1.051 0.895 0.574 0.651 1.589 1.398 1.160 0.981 0.803 0.427 0.739 1.602 1.373 1.117 0.908 0.703 0.262 0.832 1.613 1.349 1.082 0.854 0.635 0.154 0.924 1.629 1.330 1.041 0.805 0.573 0.051 1.000 1.636 1.313 1.014 0.765 0.540 0.000 Table 3. Normalized Position and Modified Scratch Displacement Data. APPENDIX C DISPLACEMENT, STRAIN, AND STRAIN-RATE HISTORIES 105 25°C, 31.1 MPa (Sample 052) 25°C, 31.1 MPa (Sample 052) Composite Sn-Ag Composite Sn-Ag Scratch Displacement Due to Creep 0 4 Localized Strain vs. Time 0.35 I I I I I I 1 I I I ' I I I I I I I I : a A I ‘ _ I . / lion-Illaui'l ...F-l' 0419 E 0 3 Elapsed 0.3 . Time ; gozs , (s...) 0.2 «I . 7 - 0.2 «1- _.._o 3 ’ — a— 240 0.1 O 2880 .5 o 15 . :9: v 4860 g o ' ‘ :g- 0.1 ’0’" 3:: 2 0.05 . - 10380 '0'1 ~- — 10920 0- ' .02 o 0.2 0.4 0.6 0.8 1 o Normalized Position Across Solder Joint Thickness (a) (b) ‘ 0.0004 4x104 . 41‘104 0.0003 g 0 0002 2x104 ’ 2"o4 4 I _ - 0.0001 Strain " » .3. Rate "L. 0 2:: 0 (US) ’\ 00001 -2x10'4 “>’\\>\ -2x10“ ' ‘ "% II 00002 .4x10“ “‘M l -4x10" -0.0003 ‘Q 4 -0.0004 -6x10“‘ ” ” ¢ \\ 6x104 ’ ’ < -0.0005 1 ” Q 10540 ’ -0.0006 0 5 ‘ 6540 - < 2540 Time ”3335113" 0 (“m“) (C) Figure 32. Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 25°C. 106 25°C, 28.9 MPa (Sample 053) 25°C, 28.9 MP3 (Sample 053) Composite Sn-Ag Composite Sn-Ag Localized Strain vs. Time Scratch Displacement Due to Creep 2 5 1 157 . ' ‘ ' ' ' ' i. ' ' ~ I 02 Normalized I I i ‘0 I .' : __E,__ 0:4 Position g/‘(g H 0.8 n_ Elapsed 2 : 06 .Through _ ”.711 ..... 3 5 Time - . 0’8 Solder Jomt // 5 g ‘ -i 0.9 Thickness ,/ i EDS (Sec.) L gjzi 1'0 : : : " g 1'5 Z Global """ Q — . . l) 0.4 c — ‘ o E 1 .......... 'O u .2 0.2 ‘0 :1: 3 0.5 --------------------- E 0 .......................... _0-2 I l i l 1 1 i l 1 l i l l 0 I v v O O O O ‘70 O O O 0 0.2 0.4 0.6 0.8 1 o 8 8 8 8 ‘— 5 v (0 Normalized Position Across N " ‘° °° ‘— 1— ‘— ‘— Solder Joint Thickness Time (sec) (a) (b) 0.00035 0.0003 3x10-4 , , 3x104 0.00025 Sn . 2x10“1 , 2x104 0.0002 am . , Rate 1 (1 /S) 0.00015 1x104 ’ , 1x104 0.0001 0.00005 0 Normalized Position Figure 33. Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 25°C. 107 25°C, 25.8 MPa (Sample 054) Composite Sn-Ag ()Sgratch Displacement Due to Creep Modified Displacement .0 9 .o .o .o 9 . N (a & OI _\ Empsed Time (KSec.) —D—o -B— 207.6 9441 v 1244 w‘ 1478 —t— 1590 r7 1753 1 0 0.2 0.4 0.6 0.8 1 Normalized Position Across Solder Joint Thickness 2x105” a IMO” I Strain Rate 0 ‘ (l/s) ‘1 -lx105’ 1 Normalized 0_ 5 Position Figure 34. 25°C, 25.8 MPa (Sample 054) Composite Sn-Ag +0 -------- 0.2 03 V 10 Global Localized Strain vs. Time Normalized 4 Position , 1 Through , j Solder Joint 3 Thickness ****** § :8: 2000 Time (ksec) O 0 ID >. sift“ 2x105 1x105 0 -1x105 1634 1034 o (Kiloseconds) (C) 108 (b) 20000 1 5000 1 0000 i 5000 -5000 -1 0000 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 25°C. 25°C, 21.9 MPa (Sample 040) 25°C, 21.9 MP3 (Sample 040) Non-Composite Sn-Ag Non-Composite Sn-Ag 1Sgratch Displacement Due to Creep 5 Localized Strain vs. Time . I l l I l l l I l I l I | l I I l | l '_ 1 . I : E i 4 f E 5 2 —: g 1-2 i : i f 3 : .1. » 3 — -------- 1 --------- l ........ ' C ............. _ g C E 1' : 3 0.8 ------- c 2 r ; , i Q ._ I I 3 I -o 0.6 E ; .; a: H 1 , .. I 5.3 . . U) 2 3/1. - "' I '80'4 --------------------- 3% —u—o1 ~o 0.8 D 3 5 0 .— —B— 0.4 + 0.9 'i 0-2 """"""" : 0 5 . 1.0 : r V 0.7 Global _,_,_‘ 0 -1 o o o o o 0 0.2 0.4 0.6 0.8 1 o 8 8 9, 8 8 Normalized Position Across ‘— " N “' Solder Joint Thickness Time (sec) (a) 0)) 0.0025 . 2 0.002 x 00.002 0 00 0.0015 , -0001 i . 0.001 . 0.001 Strain Rate (Us) 0.0005 0 r , 0 -o 0005’ , 00005 0 1 Normalized '0-0005 Position ' Time (C) Figure 35. Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 25°C. 109 25°C, 19.9 MPa (Sample 038) Non-Composite Sn-Ag Sgratch Displacement Due to Creep (I‘illIIllIllIiIlI Elapsed Time (Sec) 1560 2820 4680 6660 7860 ' ' l r' 8520 Q'.-{. Modified Displacement § o r._1. @7g14@.$4 0 0.2 0.4 0.6 0.8 1 Normalized Position Across Solder Joint Thickness (a) ##I—LJ Normalized Position Figure 36. 25°C, 19.9 MPa (Sample 038) Composite Sn-Ag chalizjegcl 1 Normalized 1 Position ' Through Solder Joint Thickness Histories for Non-Composite Solder at 25°C. 6000 8000 Strain vs. Time 88 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) 25°C, 19.8 MPa (Sample 039) 25°C, 19.8 MPa (Sample 039) Non-Composite Sn-Ag Composite Sn-Ag 1Sgratch Displacement Due to Creep 4 Localized Strain vs. Time _ FlllIlllIlllIllIIIl’T_ E 1 E El d 3 apse 8 0-8 _ Time <0 3- (366.) C 2 .2 0-6 '8 Q h .5 m .9 0.4 1 t: 1: 2° 02 . 0 , 0 ”Lillililililllii‘ 0 0.2 0.4 0.6 0.8 1 o 8 8 8 8 Normalized Position Across N v ‘° °° Solder Joint Thickness Time (sec) (60 (b) 1‘ / '\E /j’ ’ X/ i > r" \\ fl 8x104’ > ’ z8x104 / \ / /‘ “7 / \ ' 4x104’ / / Strain , 5 Rate (Us) 0 ’0 -2x10“‘ , '2x‘04 1 8240 0.5 5240 Time Normalized 0 2240 (Seconds) Position (C) Figure 37. Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 25°C. 111 25°C, 10.5 MPa (Sample 009) 25°C, 10.5 MPa (Sample 009) Non-Composite Sn-Ag Non-Composite Sn-Ag 0Sgratch Displacement Due to Creep 0 4 Localized Strain vs. Time _ I l I I I I I I I I I I I I I I I I I - _‘ C 2 11.1 _ E . 2 0 _ U _ E - Q .4 .9 r O _— 'U _ .92 - t _. 'U ' ‘ 0 fl 5 Lu» : -o‘1 0 1 I L l J i l l l l 1 l l l 1 l 1 l 1 i l l l | — 0 0.2 0.4 0.6 0.8 1 o “:9 “’9 ”$3 ”52 .08 Normalized Position Across ‘- N N v “7 Solder Joint Thickness Tlme (590) (a) (b) 0.00000012 7, I 7 12X10 12X10 0.0000001 8x108” ’ 8x108 3., 0.00000008 0.00000006 4xl0's ’ ' 4"“) 8 0.00000004 Strain 000000002 Rate 0 / , 0 (Us) 0 -4x10'8 / —-4x10"‘ 000000002 4377 . -0.00000004 Normalized POSi‘lO" 1377 Time Figure 38. Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non—Composite Solder at 25°C. 112 25°C, 11 MPa (Sample 042) 25°C, 11 MPa (Sample 042) Non-Composite Sn-Ag Non-Composite Sn-Ag Scratch Displacement Due to Creep o 3 Localized Strain vs. Time 0.7 I I I I I I I I I I I I l' I I I T u 0 Normalized 0 19 a _B__ 0.3 Posmon / .. 0.6 . 0-25 A 08 Across , )L 5 v 09 Solder Joint : ,x; E 0.5 0.2 r 1 0 Thickness j s . 3 a ............... . __. _ 0.4 0.15 _ .3 7': 2 90.3 ..0.1 “i g a :a : 5'. I?» L :3 0-2 0.05 _ _. . , : o : : 5 0.1 0 7 ¥(:2 """"" 1 """"""" . """ 1 o -o.05 I—IlllIllllIll l'I11_.l_LIllllIJ_lllA o 0.2 0.4 0.6 0.8 1 o °2 '9 ”$2 “’2 “’2 “’2 Normalized Position Across " N _ m " "’ ‘° Solder Joint Thickness Time (890) (a) (b) m 0% 'mo 7 0.0000002 0.00000015 1x10'7‘ -1x10'7 0.0000001 Strain 000000005 Rate 0 ” 0 0 (Us) I, -1x10'7‘ “1x104 -0.00000005 ‘ 00000001 -2x10 7‘ 2-2x10 7 000000015 -0.0000002 .7_ _ _7 '3‘”) '3X10 000000025 I 00000003 - 0.5 2x103 Normalized Time Posmon (Kiloseconds) (C) Figure 39. Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 25°C. 113 85°C, 10.8 MPa (Sample 033) Composite Sn-Ag 0Sgratch Displacement Due to Creep . IIIIIIIIIIIIIIIIIII" Li 0.7 """""""""" r """ 2,2 _, ; Elapsed E .- ' ‘ Time 80.5 7‘ —D—0 5! ,1 .2 —a— 120 30.4 r “r“- ';I-~g y‘_ 420 5 2 Of ’7 .-’f I ' 1 v 950 130.3 -.’- 9". - o 1680 3.3 . ' " ; —a2- 2100 502 — ..... ,2 . 2700 2° I . ,0 . I , .n 3240 0.1 ‘ ' r j , 1 gear, ‘70- : e 4620 o £4.‘9‘_I;L_L_L_L_J_1_LI__LL_1_I_LL_J_3 0 0.2 0.4 0.6 0.8 1 Normalized Position Across Solder Joint Thickness (a) 6xlO‘4 4x104 Strain Rate 4 (“5) 2X10 -2x104 1 Normalized 05 Position (C) Figure 40. 85°C, 10.8 MPa (Sample 033b) Composite Sn-Ag --------10 ‘ l . l m; l. 1 r J 51.4 smlefile¥LJ_LLi_L 0.5 - 5'. 5 Global E . ° fol/0‘0 a I i 0.46 f e _o.5 TILLIIJLLJI1114I111L11111 Time (sec) 09) 5. = 0.0007 5 0.0006 6x10“1 0.0005 4x104 0.0004 0.0003 .4 2"”) 0.0002 0 0.0001 0 .4 '2‘” -0.0001 4820 -0.0002 3120 1820 Time (Seconds) Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 85°C. 114 85°C, 10.4 MPa (Sample 0345) 85°C. 10.4 MPa (Sample 0345) Composite Sn-Ag Composite Sn-Ag Localized Strain vs. Time I l I I I I l I I I I I I I I I "I if I g070 3 59 S1cratch Displacement Due to Creep IIIITIII'IIIII Modified Displacement o 0.2 0.4 0.6 0.8 1 Normalized Position Across Solder Joint Thickness (a) , (b) 0.003 0003 0002 0.002 2 0.001 0.001 Strain I, Rate 0’ '0 ( l/s) - 0 2-0001 -0.001 '-0002 - -0.003 -0-002 1680 Normalized '0'003 Position 0 (Seconds) (C) Figure 41. Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 85°C. 115 85°C, 12.3 MPa (Sample 018b) Non-Composite Sn-Ag 2Sfcratch Displacement Due to Creep Modified Displacement l I IIIIIIIIIIII III] Elapsed Time (Sec) —D—0 -El— 360 780 ——t-— 1800 O , 1920 — - I-— 1980 0 0.2 0.4 0.6 0.8 1 Normalized Position Through Solder Joint Thickness Figure 42. Normalized Position A 32 l\ \ \//X \\\ \ ‘KVIA-V..- \\\ . wart-Ilse (C) 85°C, 12.3 MPa (Sample 018b) Non-Composite Sn-Ag Localized Strain vs. Time 3 I 2.5 2 1.5 E .E : g 1 :— ------------------ ‘D E 0.5 E. 0 —0.5 :— _1 :ALI l l I l l I l I | l 1 I I I l l | O O C) O 0 g; 8 8 8 1— ‘— N Time (see) (b) (Seconds) $20005 20.004 , 0.003 Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 85°C. 116 85°C, 16.3 MPa (Sample 025) Non-Composite Sn-Ag 0Sagratch Displacement Due to Creep _ IIIIIIIIIW’TITIIIWI p placement Modified Dis 0 Figure 43. 0.2 0.4 0.6 0.8 1 Normalized Position Through Solder Joint Thickness Normalized Position \ r A\ V ‘VA‘IIA'! 1'1‘I1‘\\\Ii \I 11““ ‘\\\ 'I” ‘I‘. i" ‘-‘"I-\-L‘\\V.\\ 'C III-I i I. .I'AII' i (C) 85°C, 17.3 MPa (Sample 025) Non-Composite Sn-Ag Localized Strain vs. Time '"Tfifi—‘T—I—T‘ WIfi—r—T—I—I—F‘TT—fi‘fiI A (Seconds) Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 85°C. 117 85°C, 16.3 MPa (Sample 026) Non-Composite Sn-Ag Scratch Displacement Due to Creep 0-5 W i 1 W71 .0 5 Modified Displacement .0 N 0 0 0.2 1.2 La_ 1 ‘9— 03 2. _,_ 0.6 0.4 0.6 0.8 1 § § Normalized Position Through Solder Joint Thickness Normalized Through / Thickness // 0555,13; : -' :1 . 1‘13,..:'; ........ .. .0 SolderJoint gf/ j U3 Time (sec) Position .. . j/'_ 85°C, 16.3 MPa (Sample 026) Non-Composite Sn-Ag Localized Strain vs. ije (a) (b) >e 0.012’< \ F0012 / A < p 0.01 / 0.008 1 \ \2 0.003 / A \ \2 0.006 0.004 ’/ \ K, 0.004 Strain / \, Rate I 0.002 (Us) 0 '0 -0.02 ‘ ”0.002 1 430 Normalized 5 330 Position ' 230 Time (Seconds) 0 (C) Figure 44. Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 85°C. 118 125°C, 9.4 MPa (Sample 028) 125°C, 9.4 MPa (Sample 028) Composite Sn-Ag Composite Sn-Ag (Spratch Displacement Due to Creep 1 Localized Strain vs. Time . I I I I | l l I I l I I l I I l I I a 0.6 0.8 C o E 0.5 § 0.6 —_ 7g“ .s - : ._ (6 . ’ _ _ 0 Q3 5 0.4 _ ; . , ' _ 1:1 . .2 : . ~ u— ........................ , 2 , , _ . i . . A :6 0 2 0.2 , 1’): ‘ ii . : I _ g 7,1? I L 7 0.1 I , . . _ 1 j I . .. 03° E E 0 0"”5" :' """" . "T OI'IIOIIIIIIIIIJ_I_I_1|1' lllilliIiiIIlllIlllIlJi o 0.2 0.4 0.6 0.8 1 o 8 8 8 8 § § . . . N v to co ‘_ ‘_ Normalized POSltlon Across ' Solder Joint Thickness Time (sec) (a) (b) 0.005 0004-” ,0_004 0.004 0.003 0.002— - 0.002 0.002 Strain Rate 0 < 0 0.001 (l/s) ’ ' 0 -0 002’ _. _0_002 -0.001 -0.002 Normalized -0.003 Position 320 Time 0 (Seconds) (C) Figure 45. Displacement (a), Localized Creep Strain (b), and Strain—Rate (c) Histories for Composite Solder at 125°C. 119 125°C, 9.9 MPa (Sample 030) Composite Sn-Ag {Spratch Displacement Due to Creep - E . . I’D—F _| a E g 7 El d apse 3 0‘8 Time 2 (See) a D — a— 2340 '5 2820 é’ 0'4 ,5 3300 E , - ~- > . 3840 20.2 . {2 —:—4°2° i I. - 1 ; e- 4140 0 :{LLLAIA l ‘._._._;_..I.1#,I I—ATTfii 0 0.2 0.4 0.6 0.8 1 Normalized Position Across Solder Joint THickness (a) 12x10 3’ ‘ > 4 \“‘ Normalized Position Figure 46. 125°C, 9.9 MPa (Sample 030) Composite Sn-Ag 1 Localized Strain vs. Time , WWFW ‘ _ 0 Normalized I. 1 -3— 03 Position . 0:4 Across . 05 Solder Joint _' i 0.8 4:2- 2 0.7 Thickness """ i 1 " 4 + 0.9 ' ' ‘ Global (b) Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Composite Solder at 125°C. 120 125°C, 11 MPa (Sample 036) Composite Sn-Ag 125°C, 11 MPa (Sample 036) Composite Sn-Ag Localized Strain vs. Time S1c2ratch Displacement Due to Creep ' ' _ ' I I I i I I I 1' Normalized ; - Position Across l l | Global Solder Joint ..l..j ..L. Thickness |1||Il | 1 i l _.D—0 I. --------- 0.3 E 1 . . Elapsed 15 7 ' g _ . . : HT Time ' 80's --::— ----- - C3, ----- (390-) : E - , . 3 \ ‘k — —u—0 1: 9- _\ —E— 1740 ._ £06 — E a? E i \t‘ 2220 g 1, Z 5 5 RR: v 2520 w a, O 4 __ _ - _ _____ L0 A ______ X "WHO”... .. 2880 SE ' - ' * o - \—' —4—3120 0. ‘8 i ‘0‘ i 2072 3420 50.2 —— ~I— 3600 : 3 § 5 : . 3660 - i i i i 0, e 3720 O I 1 ll II] III III 0 0 0.2 0.4 0.6 0.8 1 o 8 Normalized Positioin Across Solder Joint Thickness 0.0015 ’ Strain Rate (l/s) 0.0005 ’ Normalized Position 0000 0000 O 0') ’ -0.0005 00 0 L00 (‘0‘? Time (see) (b) Figure 47. Displacement (a), Localized Creep Strain (b), and Strain—Rate (c) Histories for Composite Solder at 125°C. 121 125°C, 8.1 MP8 (Sample 037) 125°C, 8.1 MPa (Sample 037) Composite Sn-Ag Composite Sn-Ag Sgratch Displacement Due to Creep 2 5 Localized Strain vs. Tlme — I I I I ~ - ..- , _ Elapsed 2 c _ II I . . I - Time ....... 8 —D—O - . 2 —E— 600 1.5 . “E """" " I n. .. 1080 E 3 ' g 1 , - v , , 1500 E .0 : 2% 1920 a 1 q, : —t— 2520 ‘ - . E ; , ,, 2940 : ' " 4 5 ‘8 0.5 """""""""""""""" g . - g— 3350 0 5 3 _ : : E It ' 4020 ' — ' g r \U 3 ~--o~~474o I . 5 § 1101 i I I 1 5700 — i l l l I i l I J 1 l l l i [ll l i l ‘ o ‘ I_l l l J i l l l i l l l I o l l A I l i 0 0.2 0.4 0.6 0.8 1 0 § § § § § § Normalized Position Across " N m " ‘° ‘° Solder Joint Thickness Time (see) (a) (b) 0.0016- EE 0.0016 0.0016 ‘ i 0.0014 0.00l2- \"0'0012 0.0012 _. \— 0.001 0.0008 - \— 0.0008 I \ 0.0008 0.0006 0~°°°4‘ ‘ r0.0004 5"“ 0.0004 R , (13:0 0.0002 0 -0 0004 —-0.0004 -0.0002 Normalized . Position 0 2080 Tim -0.0004 (Seconds) (C) Figure 48. Displacement (a), Localized Creep Strain 0)), and Strain-Rate (c) Histories for Composite Solder at 125°C. 122 125°C, 9.7 MPa (Sample 024) 125°C, 9.7 MPa (Sample 024) Non-Composite Sn-Ag Non-Composite Sn-Ag (Sgratch Displacement Due to Creep 1 2 Localized Strain vs. Time ' - - . 'i——. '_5 j"? ‘ D_ Normalized . if?» 5:0,), f/ EIapIsed _ B— 8 1 POSlthl‘l A H I '7 ,, "I l " '5’ ‘ Tlme 1 02 Across ___.r:,._|:l__ 5 0.4 """"" (Sec) V 0 3 Solder Joint I E —o— o o 0 5 Thickness - 3 —El— 420 0'8 —-— 1.0 """ F """ H ’T ¢_u 0.3 -------------- 720 ‘ ,1 _ a 1200 1: g 1’ 1440 g 0‘6 — g 02 ____________________ —-— 1680 5 j a, 2040 0 4 _ E l— 2340 _ 1:: 2530 : £04 """"""""""""""""" e~~2640 02 —.— 3000 3 . . ~ 3240 0 1 l l i l I l i l l l i 1 1 I 0 ' o o o o o o o 0 0.2 0.4 0.6 0.8 1 o 8 8 8 8 8 § § Normalized Position Across ‘— " N N Solder Joint Thickness Tlme (sec) (a) (b) 0.0012 ’ ' 0.0012 0.0008 ’ \, 0.0008 0.0004 ‘ , 0.0004 Strain Rate (“8) 0 , -0 —0 0002 - L-0.0002 Normalized Position 920 Time 0 (Seconds) (c) Figure 49. Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 125°C. 123 125°C, 10.9 MPa (Sample 027) 125°C, 10.9 MP3 (Sample 027) Non-Composite Sn-Ag Non-Composite Sn-Ag Scratch Dplaceent Due to Creep 1 4 Localized Strain vs. Time ’9/li' ’D—— Jig-=5" . I " 1.2 . '0'77708 g; 20.8 -- ~ 0.2 --I— 0.9 - 3 a: lo ; a E 1 Global " 1 8 . .4 20-6 0.8 ,_‘1 0. c _( .2 '5 I ‘ a “0.6 L’ 130.4 5 A d) I g 0.4 -/ _ o ,2 ~~~~~~~~~~~~~~~~~~~~ . ------- 5” 50 ‘- ’9' 3 2 3 0.2 x 1 E E «n ‘27:- .' 3 3 3.: Orllllilllilllilll o lllilllIiIlllilLllillllillll'ila o 0.2 0.4 0.6 0.8 1 0 § 3 § § 3 § Normalized Position Across ‘— " N N ‘° Solder Joint Thickness Time (sec) (a) (b) 0.0025- ’ 7), 0.0025 0,0025 0.002 0.0015' \, 0.0015 0.0015 Strain Rate 0.001 (l/s) , 0 0005 - 0 0.0005 -0 0005' — -0 0005 0 3080 Normalized Position Time "0‘0005 0 (Seconds) Figure 50. Displacement (a), Localized Creep Strain (b), and Strain-Rate (c) Histories for Non-Composite Solder at 125°C. 124 APPENDIX D GLOBAL SECONDARY STRAIN-RATE DATA 125 25°C, 31.1 MPa, Composite Sn-Ag (Sample 052) Global Strain History l ' l ' l ' I ' 0.35 l I l l l r r —e— Global Strain I l l l l l l l . I 0.3 0.25 0.2 0.15 Strain 0.1 0.05 Ill1lllllllllllllTITllllllll o 2000 4000 6000 8000 1 104 1.2104 Time (sec) Figure 51. Global Secondary Strain Rate for 25°C, 31.1 MPa, Composite Solder 25°C, 28.9 MPa, Composite Sn-Ag (Sample 053) Global Strain History .1...!l,.,ll 1 T T l I l l l I l I ~ —e—Global Strain f § § 5 g 5 _ Strain 0 2000 4000 6000 8000 1104 1.2104 1.4104 1.6104 Time (sec) Figure 52. Global Secondary Strain Rate for 25°C, 28.9 MPa, Composite Solder 126 25°C, 25.8 MPa, Composite Sn-Ag (Sample 054) 0 3 Global Strain History , , , , ,,,,,, I ,,,,,,,, fl , , , T l i T T ‘Y T i Y T I T Figure 53. Global Secondary Strain Rate for 25°C, 25.8 MPa, Composite Solder 25°C, 21.9 MPa, Non-Composite Sn-Ag (Sample 040) Global Strain History T I I , , , , 1 , Strain l .......................................................................................................... ' Q ’— l I : a ’- I I I 0 2 l 1 l 1 l 1 l 4 1 i l l l l i 1 1 l l i l 1 O Time (sec) Figure 54. Global Secondary Strain Rate for 25°C, 21.9 MPa, Non-Composite Solder 127 25°C, 19.9 MPa, Non-Composite Sn-Ag (Sample 038) Global Strain History . T r fl , I T . . l . I . q T 1.6 l r I l Strain Time (sec) Figure 55. Global Secondary Strain Rate for 25°C, 19.9 MPa, Non-Composite Solder 25°C, 19.8 MPa, Non-Composite Sn-Ag (Sample 039) Global Strain History , I fl , , I l , a : , , 009 T V T I l T 0.8 —9— Strain _ .E E U) o 2000 4000 6000 8000 1 10‘ Time (sec) Figure 56 Global Secondary Strain Rate for 25°C, 19.8 MPa, Non-Composite Solder 128 25°C, 10.5 MPa, Non-Composite Sn-Ag (Sample 009) Global Strain History r . I I a T a l . 0_3 i T i l I T r 0.25 0.2 0.15 Strain 0.1 0.05 ill llTl llll [ill llll llll ‘T‘ l l l l l _0 05 1 1 l 1 i l 1 1 1 i J 1 m 1 i 1 I l 1 i l l 1 l O —L _x O a: N .1. O C» 00 _L O a) A _x O o: 01 _L O a) Time (sec) Figure 57. Global Secondary Strain Rate for 25°C, 10.5 MPa, Non-Composite Solder 25°C, 11 MPa, Non-Composite Sn-Ag (Sample 042) Global Strain History 0.25 l l T T J I l l r l 0.2 0.15 Strain 0.1 0.05 o 1106 2106 3106 4106 5106 6106 Time (sec) Figure 58. Global Secondary Strain Rate for 25°C, 11 MPa, Non-Composite Solder 129 85°C, 10.8 MPa, Composite Sn-Ag (Sample 033b) Global Strain History . . I 1 , a f ! . 0.8 l T a T rT —e—Strain ' l l fii i l l r J11! 0.7 0.6 WTfIllll 0.5 flfi l Strain 0.4 0.3 —1 0.2 rjlrlllilIlTTTrT 0.1 2000 3000 4000 5000 O _x O O 0 Time (sec) Figure 59. Global Secondary Strain Rate for 85°C, 10.8 MPa, Composite Solder 85°C, 10.4 MPa, Composite Sn-Ag (Sample 034b) Global Strain History l T l I l i 17 l i T T l 0.7 l l l i i 1 l lllll Strain O 500 1000 1500 2000 Time (sec) Figure 60. Global Secondary Strain Rate for 85°C, 10.4 MPa, Composite Solder 130 85°C, 12.3 MPa, Non-Composite Sn-Ag (Sample 018b) Global Strain History a r , I Y , , I I T T 2.5 f l T I l r l +suam § 3 g ( Strain 0 500 1000 1 500 2000 Time (sec) Figure 61. Global Secondary Strain Rate for 85°C, 12.3 MPa, Non-Composite Solder 85°C, 17.3 MPa, Non-Composite Sn-Ag (Sample 025) Global Strain History , T I , , . , 1 1 0.35 , T T 1 , —e—Straln ‘ ............................................................................................................ .................................................................................................................. ........................................................................................................ Time (Sec) Figure 62. Global Secondary Strain Rate for 85°C, 17.3 MPa, Non-Composite Solder 131 85°C, 16.3 MPa, Non-Composite Sn-Ag (Sample 026) Glob 0.5 r l T i l I l l ' l l l I l l l l I l aI Strain History 1 l ' l ' l l l I. 17 l ‘ —e— Strain .E Q 5 100 150 200 250 300 350 400 450 Time (sec) Figure 63. Global Secondary Strain Rate for 85°C, 16.3 MPa, Non-Composite Solder 125°C, 9.4 MPa, Composite Sn-Ag (Sample 028) Global Strain History 0.5 T 1 T i , I . 7 , . , I , , 1‘ r 4 -—e— Strain ' : .s _: g T ‘ _ 5 f ............................................... _‘ 0 : l i L 1 i 1 i 1 i i l 1 L l : 0 200 400 600 800 1000 1200 Time (sec) Figure 64. Global Secondary Strain Rate for 125°C, 9.4 MPa, Composite Solder 132 Strain Figure 65. Strain Figure 66. 1.2 0.8 _ 0.6 0.4 0.2 500 1.2 0.8 0.6 0.4 0.2 125°C, 9.9 MPa, Composite Sn-Ag (Sample 030) Global Strain History i l i I I I I l I I i I l I I I i i I i I I I I 1 ' l ‘ —e— Strain l l l i I T l l i I T i l T I ............................................................................................................ .................................................................... : § j is = 1.7'7e-04 l lIlll4lllillli|iiJLil llillliillll lei 2000 2500 3000 3500 4000 Time (sec) 1000 1500 Global Secondary Strain Rate for 125°C, 9.9 MPa, Composite Solder 125°C, 11 MPa, Composite Sn-Ag (Sample 036) Global Strain History 4500 l T —e— Strain I 7 l T i l l I i l 7 l I l i I T I I T l i l l l l I T l T l I T l T | ...................................................................................................... .............................................................................................. ................................................................................................................... i I l i l I l l i l l l 1 i i 1 l l l i l l I i l l l l l i l l 1 CL 2000 2500 3000 3500 Time (sec) 500 1000 1500 Global Secondary Strain Rate for 125°C, 11 MPa, Composite Solder 133 4000 2 T 125°C, 8.1 MPa, Composite Sn-Ag (Sample 037) Global Strain History 1 ' ‘ ' ‘ l ‘ ' ' ‘ T T j T l T T I l T T T I l l —e— Strain .s 92 5 0 1000 2000 3000 4000 5000 6000 Time (sec) Figure 67. Global Secondary Strain Rate for 125°C, 8.1 MPa, Composite Solder 125°C, 9.7 MPa, Non-Composite Sn-Ag (Sample 024) Global Strain History 0.5 I ' I i T I T l ’- 1 I l T l l I T l T l I l T T l I T i T i _I T T T j A j —e— Straln ' ‘ 0.4 - ------------------------------------------------------ 03 ; .................................................................................. .5 t E _ (D 0.2 L 01 f .. . ............. O :_ 1 I i L 1 1 i..._1___1_ 1. 1 i l 1 1 1 i_1_..1___1_.___1__i 1 l 1 l '1 l 1 1 .1- ~ 0 500 1000 1500 2000 2500 3000 3500 Time (sec) Figure 68. Global Secondary Strain Rate for 125°C, 9.7 MPa, Non-Composite Solder 134 125°C, 10.9 MPa, Non-Composite Sn-Ag (Sample 027) Global Strain History W,,.. 1 l i l l I I I l l I . -9-— Strain 3 g i : _ l i I T r ‘r T I T T T V Strain ........................................................................................... 0 500 1000 1500 2000 2500 3000 3500 Time (sec) Figure 69. Global Secondary Strain Rate for 125°C, 31.1 MPa, Composite Solder 135 APPENDIX E FRACTURE SURFACES AND PROFILES 136 Stress OGTC % Sample Type °C (MPa) 78 (3'1) OGTC OLTC OLTC Porosity 052 Composite 25 31.1 1.33E-05 0.16 0.19 0.84 0.09 Figure 70. Fracture Surfaces and Profile for 25°C, 31.1 MPa, Composite Solder. 137 Stress OGTC % ‘Sample Type °C (MPa) is (5'1) OGTC OLTC OLTC Porosity 053 Composite 25 28.9 4.15E-05 0.52 1.2 0.43 0.025 Figure 71. Fracture Surfaces and Profile for 25°C, 28.9 MPa, Composite Solder. 138 Stress OGTC % Sample Type °c (MPa) ,5 E1) OGTC OLTC OLTC Porosity | 040 Non-Compos'_ite 25 21.9 4.85E—04 0.9 3 0.3 0.13 Figure 72. Fracture Surfaces and Profile for 25°C, 21.9 MPa, Non-Composite Solder. 139 ~ ill .i‘l.\ , ~ILI OGTC OLTC OLTC Porosi s") ( Figure 73. Fracture Surfaces and Profile for 25°C, 19.9 MPa, Non-Composite Solder. 140 Stress OGTC % iSample Type °c (MPa) 75(s'1) OGTC OLTC 0ch Porosity 039 Non-Composite 25 19.8 8.42E-05 0.79 3 0.26 0.064 Figure 74. Fracture Surfaces and Profile for 25°C, 19.8 MPa, Non-Composite Solder. 141 ‘0 yr“, ,. ca 1 i A" . Sample Type 71. (8'1) OLTC Porosity I 009 Non-Composite 7.13E-08 0.25 0.02 Figure 75. Fracture Surfaces and Profile for 25°C, 10.5 MPa, Non-Composite Solder. 142 Stress OGTC % Sample Type °C (MPa) 7-5 (s'1) OGTC OLTC OLTC Porosity 033 Composite 85 10.8 9.75E-05 0.5 0.9 0.56 0.21 Figure 76. Fracture Surfaces and Profile for 85°C, 10.8 MPa, Composite Solder. 143 OGTC OLTC TC Porosity Figure 77. Fracture Surfaces and Profile for 85°C, 10.4 MPa, Composite Solder. 144 Stress OGTC % Sample Type °c (MPa) ,5 (s‘) OGTC OLTC OLTC Porosity 018 Non-Compositg 85 12.3 2.82E-04 1.3 1.67 0.78 0.036 Figure 78. Fracture Surfaces and Profile for 85°C, 12.3 MPa, Non-Composite Solder. 145 Stress OGTC % Sample Type °c (MPa) ,5 (s‘) OGTC OLTC OLTC Porosity [ 025 Non-Composite 85 17.3 1.24E-04 0.24 0.3 0.8 0.11 Figure 79. Fracture Surfaces and Profile for 85°C, 17.3 MPa, Non-Composite Solder. 146 Stress OGTC % Sample Type °c (MPa) ,5 (s4) OGTC OLTC OLTC Porosity | 026 Non-Composite 85 16.3 8.4E-04 0.22 0.5 0.4 0.043 Figure 80. Fracture Surfaces and Profile for 85°C, 16.3 MPa, Non-Composite Solder. 147 Stress OGTC % Sample Type °c (MPa) ,5 (3'1) OGTC OLTC OLTC Porosity 028 Composite 125 9.4 3.92E-04 0.28 0.3 0.93 0.24 Figure 81. Fracture Surfaces and Profile for 125°C, 9.4 MPa, Composite Solder. 148 Stress OGTC % Sample Type °c (MPa) ,5 (s4) OGTC OLTC OLTC Porosity 030 Composite 125 9.9 1.77E-04 0.3 0.3 1 0.19 Figure 82. Fracture Surfaces and Profile for 125°C, 9.9 MPa, Composite Solder. 149 Stress OGTC % Sample Type °C (MPa) ,5 (3'1) OGTC OLTC OLTC Porosity 036 Composite 125 11 2.26E-04 0.11 0.15 0.73 0.14 Figure 83. Fracture Surfaces and Profile for 125°C, 11 MPa, Composite Solder. 150 .‘4 with. 'l" *5 *w Jr. -{' i "r . «rut 1 l ";‘-d.. Stress OGTC % Sample Type °C (MPa) ,5 (3'1) OGTC OLTC OLTC Porosity 037 Composite 125 8.1 2.27E-04 1.1 1.29 0.85 0.024 Figure 84. Fracture Surfaces and Profile for 125°C, 8.1 MPa, Composite Solder. 151 ‘ . 5 _t-rr_ \\E' . _-'IIIIF_ ' q Stress OGTC % Sample Type °C (MPa) ,5 (S1) OGTC OLTC OLTC Porosity L 024 Non-Cothe125 9.7 1.0E—04 0.24 0.64 0.38 0.046 Figure 85. Fracture Surfaces and Profile for 125°C, 9.7 MPa, Non-Composite Solder. 152 OGTC OLTC Porosity Figure 86. Fracture Surfaces and Profile for 125°C, 10.9 MPa, Non-Composite Solder. 153 APPENDIX F SOLDER JOINT CREEP DATA 154 Stress Secondary OGTC % Porosity of Sample Type °C (MPa) Creep-Rate OGTC OLTC OLTC 1333““ ll ace 052 C 25 31.1 1.33E-05 0.16 0.19 0.84 0.091 053 C 25 28.9 4.15E-05 0.52 1.2 0.43 0.025 054 C 25 25.8 1.28E-07 0.14 0.3 0.47 NA 040 NC 25 21.9 4.85E-04 0.9 3 0.30 0.128 038 NC 25 19.9 8.54E-05 0.86 4 0.22 0.053 039 NC 25 19.8 8.42E-05 0.79 3 0.26 0.064 009 NC 25 10.5 7.13E-08 0.2 0.25 0.80 0.02 042 NC 25 11 3.14E-08 0.15 0.19 0.79 NA 033a C 85 7.4 3.05E-06 NA NA NA 0.205 033b C 85 10.8 9.75E-05 0.5 0.9 0.56 0.205 (+ 0.5 lb.) 034a C 85 8.1 4.71 E-06 NA NA NA 0.222 034D C 85 10.4 1.46E-04 0.4 0.7 0.57 0.222 (+ 0.7 lb.) 0183 NC 85 9.8 1.39E-05 NA NA NA 0.036 018b NC 85 12.3 2.82E-04 1.3 1.67 0.78 0.036 (+0.75 lb.) 025 NC 85 17.3 1.24E-04 0.24 0.3 0.80 0.11 026 NC 85 16.3 8.40E-04 0.22 0.55 0.40 0.043 028 C 125 9.4 3.92 E-04 0.28 0.3 0.93 0.242 030 C 125 9.9 1.77E-04 0.3 0.3 1.00 0.187 036 C 125 11 2.26E-04 0.11 0.15 0.73 0.14 037 C 125 8.1 2.27E-04 1.1 1.29 0.85 0.024 024 NC 125 9.7 1.03E-04 0.24 0.64 0.38 0.046 027 NC 125 10.9 1.37E-04 0.46 0.97 0.47 0.034 Definitions + () lb. Increase in load upon establishing a secondary creep-rate C Composite Solder NC Non-Composite Solder OGTC Onset of Global Tertiary Creg) OLTC Onset of Local Tertiary Creep % Porosity of Fracture Surface Average percent of the two fractured load bearing cross- sectional areas characterized by voids and porosity. Table 4. Solder Joint Creep Data. 155 REFERENCES 156 [1] [2] [3] l4] [5] [6] l7] [8] [9] [10] REFERENCES Vianco, Paul T. and Darrel R. Frear, “Issues in the Replacement of Lead Bearing Solders,” JOM, 1993, Vol. 45, No. 7, pp. 14-19. Samans, Carl H., Engineering Metals and Their Alloys, The MacMillan Co., New York, © 1949, pp. 858 - 864. Vianco, Paul T., “Soldering in Electronic Applications,” ASM Handbook, 1990, Vol. 6, pp. 985-1000. Manko, Howard H., Solders and Soldering - Materials, Design, Production, and Analysis for Reliable Bonding, 3'“ Ed., McGraw-Hill, Inc., New York, (0 1992,p.1. Hwang, J.S., and RM. Vargas, “Solder Joint Reliability - Can Solder Creep?” Soldering & Surface Mount Technology, No. 5, June 1990, p. 38. MacKay, Colin, “Flux Reactions and Solderability”, Solder Joint Reliability - Theory and Applications, ed. John H. Lau, Van Nostrand Reinhold, New York, @1991, pp. 1-2. Yang, Hong, Phillip Deane, Paul Magill, and K. Linga Murty, “Creep Deformation of 96.5Sn-3.5Ag Solder Joints in a Flip Chip Package”, 1996 Electronic Components and Technolog) Conference, ©1996 IEEE, p. 1136. Miller, Chad M., Iver E. Anderson, and Jack F. Smith, “A Viable Tin-Lead Solder Substitute: Sn-Ag-Cu”, Journal of Electronic Materials, Vol. 23, No. ?, 1994, p. 595. Yang, Wenge, Lawrence E. Felton, and Robert W. Messler, Jr., “The Effect of Soldering Process Variables on the Microstructure and Mechanical Properties of Eutectic Sn-Ag/Cu Solder Joints”, Journal of Electronic Materials, Vol. 24, No. 10, 1995, pp. 1465-1471. Choi, S.L., A.W. Gibson, J.L. McDougall, T.R. Bieler, K.N. Subramanian, “Mechanical Properties of Sn-Ag Composite Solder Joints Containing Copper-Based Intermetallics”, Design and Reliability of Solders and Solder Interconnections. R.K. Mahidhara, D.R. Frear, S.M.L. Sastry, K.L. Murty, P.K. Liaw, W.L. Winterbottom, eds. TMS, Warrendale, PA 1997, p. 241. 157 [11] [12] [13] [14] [15] [16] [17] [21] [22] Raeder, C.H., G.D. Schmeelk, D. Mitlin, T. Barbieri, W. Yang, L.F. Felton, R.W. Messler, D.B. Knorr, and D. Lee, “Isothermal Creep of Eutectic SnBi and SnAg Solder and Solder Joints”, 1994 IEEE/CPMT International Electronics Manufacturing Technology Symposium, pp. 1, 5. Frost, Harold J ., Robert T. Howard, Patrick R. Lavery, and Scott D. Lutender, “Creep and Tensile Behavior of Lead-Rich Lead-Tin Solder Alloys’, IEEE Transactions on Components, Hybrids, and Manufacturing Technology, Vol. 11, No. 4, December 1988, p. 379. Gibson, A.W., S.L. Choi, K.N. Subrarnanian, T.R. Bieler, “Issues Regarding Microstructural Coarsening Due to Aging of Eutectic Tin-Silver Solder”, Reliability of Solders and Solder Joints, TMS Conference Orlando 1997. Yang, Wenge, and Robert Messler, Jr., “Microstructural Evolution of Eutectic Sn- Ag Solder Joints”, Journal of Electronic Materials, Vol. 23, No. 8, 1994, pp. 765-772. McCormack, M., S. Jin, and G.W. Kammlott, “Supression of Microstructural Coarsening and Creep Deformation in a Lead-Free Solder”, Appl. Phys. Lett., Vol. 64, No. 5, 31 January 1994, pp. 580-582. Garofalo, Frank, Fundamentals of Creep and Creep-Rupture in Metals, The MacMillian Company, New York, © 1965, pp. 1-20. Ross, R.G., Jr., L.C. Wen, G.R. Mon, “Solder Joint Creep and Stress Relaxation Dependence on Construction and Environmental Stress Parameters”, Journal of Electronic Packaging, Vol. 115, June, 1993, pp.165-171. Evans, R.W. and B. Wilshire, Introduction to Creep, The Institute of Materials, @1993, pp. 1-31. Glen, John, The Problem of the Creep of Metals, Murex Welding Processes LTD., Waltham Cross, Hertfordshire, England, January 1968, pp. 8 - 15. Evans, R.W. and B. Wilshire, Creep of Metals and Alloys, The Institute of Metals, @1985, pp. 37-57, 70-76. Dieter, George E., Mechanical Metallurgy, 3rd Edition, McGraw-Hill Publishing Company, New York, © 1986, pp. 310-314, 438 - 449. Reed-Hill, Robert E. and Abbaschian, Physical Metallurgy Principles, 3’“ Edition, PWS-Kent Publishing Company, Boston, © 1992, p 854. 158 [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] Bird, J.E., A.K. Mukherjee, and IF. Dorn, “Correlations Between High- Temperature Creep Behavior and Structure”, Quantitative Relations Between Properties and Microstructure, Proceedings of an International Conference, Haifa, Isreal, July 27-Aug. 4, 1969, p. 255. Weertman, J ., “Dislocation Climb Theory of Steady-State Creep”, Creep Symposium, Transactions of the ASM vol. 61, ©1968. Artz, E. M.F. Ashby, R.A. Verrall, “Interface Controlled Diffusional Creep”. Acta Metall. Vol 31, #12. 1983. Ashby, M.F. “Boundary Defects, and Atomistic Aspects of Boundary Sliding and Diffilsional Creep”, Surface Science, 31, 1972. Ruano, O.A., J. Wadsworth, J. Wolfenstine, O.D. Sherby, “Evidence for Nabarro Herring creep in metals: fiction or reality?”, Materials Science and Engineering A165, 1993. Wolfenstine, J ., O.A. Ruano, J. Wadsworth, O.D. Sherby, “Refutation of the Relationship Between Denuded Zones and Diffusional Creep”, Scripta Metallurgica vol. 29, 1993. Ross, Ronald 0., Jr., and Liang-chi Wen, “Crack Propagation in Solder Joints During Thermal-Mechanical Cycling”, Journal of Electronic Packaging, June, 1994, Vol. 116, p. 69. Gibson, A.W., S. Choi, T.R. Bieler, and K.N. Subrarnanian, “Environmental Concerns and Materials Issues in Manufactured Solder Joints”, 1997 IEEE International Symposium on Electronics and the Environment, (IEEE, Piscataway, NJ, 1997), pp. 246-51. Frost, Harold J., “Microstructure and Mechanical Properties of Solder Alloys”, Solder Joint Reliability - Theory and Applications, ed. John H. Lau, Van Nostrand Reinhold, New York, ©1991 , pp. 266-267. Frear, Darrel R., “Thermomechanical Fatigue in Solder Materials”, Solder Mechanics - A State of the Art Assessment, eds. Darrel R. Frear, Wendell B. Jones, and Kenneth Kinsman, The Minerals, Metals, and Materials Society, Warrendale, PA, 1991, p. 191. Frost, Harold J ., and Robert T. Howard, “Creep Fatigue Modeling for Solder Joint Reliability Predictions Including the Microstructural Evolution of the Solder”, IEEE Transactions on Components, Hybrids, and Manufacturing Technology, Vol. 13, No. 4, Dec. 1990, p. 727. 159 [34] [35] [36] [37] [38] [39] [40] [41] [42] Attarwala, A.I., J .K. Tien, G.Y. Masada, G. Dody, “Confirmation of Creep and Fatigue Damage in Pb/Sn Solder Joints”, Journal of Electronic Packaging, June 1992, Vol. 114, pp. 109,111. ‘ Vaynman, Semyon, and Moris E. Fine, “Effects of Strain Range, Ramp Time, Hold Time, and Temperature on Isothermal Fatigue Life of Tin-Lead Solder Alloys”, Solder Joint Reliability - Theory and Applications, ed. John H. Lau, Van Nostrand Reinhold, New York, ©1991, pp. 333,357. Hall, Peter M., “Creep and Stress Relaxation in Solder Joints”, Solder Joint Reliability - Theory and Applications, ed. John H. Lau, Van Nostrand Reinhold, New York, @1991, pp. 313-332. Morris, J.W., Jr., D. Tribula, T.S.E. Summers, and D. Grivas, “The Role of Microstructure in Thermal Fatigue of Pb-Sn Solder Joints”, Solder Joint Reliability - Theory and Applications, ed. John H. Lau, Van Nostrand Reinhold, New York, ©1991, pp. 225-226, 261-262. Tien, John K., Bryan C, Hendrix, and Abbas I. Attawala, “The Interaction of Creep and Fatigue in Lead-Tin Solders”, Solder Joint Reliability - Theory and Applications, ed. John H. Lau, Van Nostrand Reinhold, New York, @1991, pp. 279-305. Dasgupta, A., C. Oyan, D. Barker, M. Pecht, “Solder Creep-Fatigue Analysis by an Energy-Partitioning Approach”, Transactions of the ASME, Journal of Electronic Packaging, Vol. 114, June, 1992, pp. 152-160. Pan, Tsung-Yu, “Critical Accumulated Strain Energy (Case) Failure Criterion for Thermal Cycling Fatigue of Solder Joints”, Journal of Electronic Packaging, Vol. 116, September, 1994, pp. 163-170. Darveaux, Robert, and Kingshuk Banerji, “Constitutive relations for Tin-Based- Solder-Joints”, I992 Preceedings of the IEEE 42"" Electronic Components & Technology Conference, pp. 538-551. Wasynczuk, J .A. and George K. Lucey, “Shear Creep of Cu6Sn5/Sn-Pb Eutectic Composites”, Proceedings of Technical Program Vol. 111, National Electronic Packaging and Production Conference, 1992, pp. 1245-1255. 160 "‘lllllllilllllllllll“