© Copyright by Shuren Qu 2019 All Rights Reserved TRIBOLOGY OF PTFE/PEEK COMPOSITE AT ELEVATED TEMPERATURE A Dissertation Presented to the Faculty of the Interdisciplinary Program in Materials Engineering University of Houston In Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in Materials Engineering by Shuren Qu August 2019 TRIBOLOGY OF PTFE/PEEK COMPOSITE AT ELEVATED TEMPERATURE Shuren Qu Approved: Chair of the Committee Su Su Wang, Hugh Roy and Lillie Cranz Cullen Professor, Department of Mechanical Engineering Committee Members: Alamgir Karim, Dow Chair and Welch Foundation Professor, Department of Chemical & Biomolecular Engineering Megan Robertson, Associate Professor, Department of Chemical & Biomolecular Engineering Akira Miyase, Research Professor, Department of Mechanical Engineering King Him Lo, Research Professor, National Wind Energy Center Suresh K. Khator, Associate Dean, Cullen College of Engineering Alamgir Karim, Dow Chair and Welch Foundation Professor and Director of the Interdisciplinary Program in Materials Engineering Acknowledgements I am sincerely thankful for all the remarkable people who had helped me, personally and professionally, through my years as a Ph.D. student at the University of Houston. I would like to thank my advisor, Professor Su Su Wang for his support and guidance throughout my research. He is not only a knowledgeable professor, but also a mentor and a friend. I have been continuously improving myself professionally as a student through his invaluable feedbacks and leadership. I also thank him for always motivating and encouraging me throughout my Ph.D. career. I would also like to express my gratitude towards Professor Alamgir Karim, Professor Megan Robertson, Professor Akira Miyase and Professor King Him Lo who served as members of my thesis committee for their effort in evaluating this work. I thank Dr. Akira Miyase and Dr. King Him Lo for training me as a new researcher at Composite Engineering and Applications Center and their significant contributions to this research. I am fortunate to work with them in the past 5 years. Their integrity, dedication and vision in research inspired me tremendously. I will certainly bring these good characters to my future endeavors. I thank Dr. Boris Makarenko from University of Houston Department of Chemistry for his kind help and fruitful discussion on the surface elemental analysis with X-ray photoelectron spectroscopy. I am also grateful to my group members, Jonathan Penaranda and Ethan Pedneau, for their endless help on my research and making graduate school as enjoyable as v possible. Thank Jonathan and Ethan for being my friends. I cherish our friendships and it is a true pleasure working with them. I would like to dedicate this work to my parents, Zhibo Qu and Jinge Tang, my grandmother, Jinhuan Gu, and my wife Ya Zhuo. I could not have finished my Ph.D. study without their love and support. Praise the Lord that I have such a wonderful family that always encourage me and give me hope. All the laughter and the tears we had together now become worthy. Thank God for carrying me so far and I will never stop my steps seeking truth. Financial assistance from the Composites Engineering and Applications Center for this research is also gratefully acknowledged. vi TRIBOLOGY OF PTFE/PEEK COMPOSITE AT ELEVATED TEMPERATURE An Abstract of a Dissertation Presented to the Faculty of the Interdisciplinary Program in Materials Engineering University of Houston In Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in Materials Engineering by Shuren Qu August 2019 vii Abstract Engineering thermoplastics are used extensively in reciprocating and rotating machinery. For example, Polyetheretherketone (PEEK) and its composites are commonly used as sealing materials in compressors and pumps. Neat PEEK polymer is known for its outstanding chemical resistance, thermal stability, and high-temperature mechanical strength. Despite these attractive properties, a critical concern of the PEEK polymer in tribological applications is its high friction, which could lead to serious problems, such as high local flash temperature and severe wear. A common practice to reduce the high friction is to incorporate solid-state lubricants, such as polytetrafluoroethylene (PTFE), into the polymer. In view of the wide use of PTFE/PEEK composite for tribological applications, predictive models and theories are needed to better understand friction and wear mechanisms of the composite and accelerate development and applications of the composite for tribological operations. In the experimental phase of this study, friction and wear of PTFE/PEEK composites with various PTFE contents were determined at different temperatures. Elevated temperature mechanical properties of neat PEEK, PTFE and their composites, such as elastic modulus and compressive yield stress, were also determined. Quantitative methods for characterization of the transfer films were developed. The area coverage ratio and composition of composite transfer films were investigated with an in-house developed computer software and XPS, respectively. Important experimental results were obtained and used for subsequent developments of friction and wear models and theories. In the theoretical phase of the study, a power-law relationship between PTFE/PEEK composite friction and wear is first established based on experimental viii results. Friction and wear models are established for developing theories on PTFE/PEEK composite sliding friction and wear at low temperature. Solid film lubrication is introduced along with the rule of the mixtures to derive the new friction and wear theories for the polymer composite. For elevated temperature friction and wear of the PTFE/PEEK composite, mechanism-based friction and wear models are proposed for the development of the new tribological theories to predict the friction and wear of PTFE/PEEK composite. Detailed mechanisms and mechanics of low and elevated temperature friction and wear of the PTFE/PEEK composites are discussed. ix Table of Contents Acknowledgements ........................................................................................................... v Abstract ................................................................................................................... viii Table of Contents .............................................................................................................. x List of Figures ................................................................................................................. xiii List of Tables ................................................................................................................ xviii Chapter 1 Introduction ................................................................................................ 1 Chapter 2 Literature Review ...................................................................................... 8 2.1 Polymer Composites with Solid-State Lubricants .................................................... 8 2.2 Friction and Wear of PTFE/PEEK Composite........................................................ 13 2.3 Effect of Transfer Films on Tribological Properties of PTFE/PEEK Composite ... 15 2.3.1 PTFE ................................................................................................................. 15 2.3.2 PEEK ................................................................................................................ 17 2.3.3 PTFE/PEEK Composite ................................................................................... 18 2.4 Experimental Methods for Polymer Tribological Study ......................................... 19 2.4.1 Block-on-Ring Tribometer ............................................................................... 19 2.4.2 Pin-on-Disk Tribometer .................................................................................... 20 2.4.3 Linear-Reciprocating Tribometer ..................................................................... 21 2.4.4 Thrust-Washer Tribometer ............................................................................... 21 2.5 Friction and Wear Theories of Polymer Composite................................................ 21 2.5.1 Friction............................................................................................................ 21 2.5.2 Transfer Films................................................................................................... 23 2.5.3 Wear.................................................................................................................. 24 2.5.4 Flash Temperature ............................................................................................ 26 Chapter 3 Objectives and Scope of Research ........................................................... 27 Chapter 4 Materials System and Experimental Program ....................................... 29 4.1 Material System....................................................................................................... 29 4.1.1 Constituent Materials ........................................................................................ 29 4.1.2 PTFE/PEEK Composite Microstructure ........................................................... 29 x 4.1.3 Polymer and Composite Morphology............................................................... 30 4.2 Thermal and Mechanical Properties ........................................................................ 34 4.2.1 Thermal Properties ........................................................................................... 34 4.2.2 High-temperature Mechanical Properties ......................................................... 35 4.3 Experimental Facilities ............................................................................................ 45 4.3.1 High-Temperature Pin-on-Disk Tribometer ..................................................... 45 4.3.2 High Temperature Stage and Temperature Controller ..................................... 48 4.4 Experimental Program............................................................................................. 50 4.4.1 Sample Preparation ........................................................................................... 50 4.4.2 Friction and Wear Test Matrix ......................................................................... 51 4.4.3 Experimental Procedure ................................................................................... 54 4.4.4 Data acquisition and analysis ........................................................................... 54 Chapter 5 Relationship between Friction and Wear of PTFE/PEEK Composite .................................................................................................................... 58 5.1 Experimental Results of Friction............................................................................. 58 5.2 Experimental Results of Wear................................................................................. 60 5.3 Relationship between Composite Friction and Wear .............................................. 62 5.4 Validation with Literature Data............................................................................... 64 Chapter 6 Transfer Films ........................................................................................... 69 6.1 Nature and Issues of Transfer films in PTFE, PEEK and Their Composite ........... 69 6.2 Experiment Methods for Transfer Film Evaluation ................................................ 70 6.2.1 Transfer Film Coverage .................................................................................... 70 6.2.2 Elemental and Compositional Analysis of Transfer Films ............................... 73 6.3 Experimental Observations ..................................................................................... 73 6.4 Characterization of Transfer Films on Counterface ................................................ 75 Chapter 7 Development of Friction and Wear Theories for PTFE/PEEK Composite .................................................................................................. 79 7.1 Friction and Mechanical Properties of PTFE/PEEK Composite ............................ 79 7.2 Solid Film Lubrication and Associated Models ...................................................... 82 7.3 Friction Involving Transfer Films ........................................................................... 84 7.4 Wear with Transfer Films ....................................................................................... 88 xi Chapter 8 Elevated Temperature Friction and Wear of PTFE/PEEK Composite .................................................................................................................... 91 8.1 Elevated Temperature Friction and Wear Experimental Results ............................ 91 8.1.1 Friction.............................................................................................................. 91 8.1.2 Wear.................................................................................................................. 93 8.2 Characteristics of Friction and Wear at Elevated Temperature .............................. 95 8.3 Relationship between Friction and Wear at Elevated Temperature ...................... 100 8.4 Elevated Temperature Friction Theory ................................................................. 103 8.4.1 Validation of Friction Theory with Room Temperature Experiments ........... 109 8.4.2 Validation of Friction Theory with Elevated Temperature Experiments ....... 112 8.5 Elevated Temperature Wear Theory ..................................................................... 113 Chapter 9 Mechanisms of Friction and Wear of PTFE/PEEK Composite at Elevated Temperature ........................................................................... 117 9.1 Friction .................................................................................................................. 117 9.1.1 Neat PEEK ...................................................................................................... 117 9.1.2 Neat PTFE ...................................................................................................... 119 9.1.3 PTFE/PEEK composite .................................................................................. 120 9.2 Wear ...................................................................................................................... 122 9.2.1 Neat PEEK ...................................................................................................... 122 9.2.2 Neat PTFE ...................................................................................................... 123 9.2.3 PTFE/PEEK Composite ................................................................................. 123 Chapter 10 Conclusions .............................................................................................. 125 References .................................................................................................................. 128 Appendix A Numerical Simulation of Randomly Distributed Spherical Particle Filled Composite ..................................................................................... 142 Appendix B Friction Coefficients by the Rule of Mixtures ...................................... 147 xii List of Figures Figure 1.1 Chemical structures of PTFE and PEEK polymers. …………… 4 Figure 2.1 Specific wear rate and friction coefficient of selected neat polymers (blue dots) and polymer composites (orange dots) sliding against a metal conterface………………………………. 9 Lamellar lattice structures of (a) graphite (b) molybdenum disulfide and (c) polymer chain of polytetrafluoroethylene…… 11 Figure 2.2 Figure 2.3 Schematic model of transfer film development of FTFE [69]….. 16 Figure 2.4 SEM pictures of steel counterface after (a) 1 cycle; (b) 10 cycles; (c) 100 cycles; (d) 10,000 cycles; (e) 70,000 cycles; and (f) 141,000 cycles sliding of PEEK…………………………….. 17 Figure 2.5 Common tribological test systems: (a) block-on-ring tribometer, (b) pin-on-disk tribometer, (c) linear reciprocating tribometer and (d) thrust washer tribometer………………………………... 20 Figure 2.6 Solid film lubrication at a polymer-steel contact junction……… 24 Figure 4.1 SEM micrographs of (a) neat PEEK, (b) C05, (c) C10, (d) C15, and (e) C20 composites, and (f) neat PTFE…………………….. 30 Figure 4.2 XRD patterns of neat PEEK, neat PTFE and PTFE/PEEK composite……………………………………………………….. 31 Crystalline and amorphous peaks of the X-ray diffraction pattern of C10…………………………………………………... 32 Figure 4.3 Figure 4.4 Crystallinity of neat PEEK, neat PTFE, and individual PEEK and PTFE phases in the PTFE/PEEK composite……………….. 34 Figure 4.5 DSC traces of neat PEEK, neat PTFE and PTFE/PEEK composites……………………………………………………… 35 Hardness of neat PEEK, neat PTFE and PTFE/PEEK composites at room temperature………………………………... 36 Figure 4.6 Figure 4.7 Storage modulus of neat PEEK, neat PTFE and PTFE/PEEK composites as a function of temperature.……………………….. 37 Figure 4.8 Loss modulus of neat PEEK, neat PTFE and PTFE/PEEK composites as a function of temperature ...……………………... 38 xiii Figure 4.9 Loss tangent of neat PEEK, neat PTFE and PTFE/PEEK composites as a function of temperature………………………... 38 Figure 4.10 Stress-strain curves of elevated temperature compression tests of (a) PEEK, (b) C05, (c) C10, (d) C15, (e) C20 and (f) PTFE… 43 Figure 4.11 Elastic moduli of neat and composite samples as a function of temperature ……………………….............................................. 44 Figure 4.12 Compressive yield stresses of neat and composite samples as a function of temperature…………………………………………. 44 Figure 4.13 Schematic of pin-on-disk tribometer…………………………… Figure 4.14 Drive train of the pin-on-disk tribometer……………………….. 46 Figure. 4.15 Assembly of load arm, load cell, and LVDT…………………… Figure 4.16 Temperature control stage for high-temperature pin-on-disk tribometer……………………………………………………….. 49 Figure 4.17 Multi-channel temperature controller and recorder…………….. 50 Figure 4.18 A steel counterface. (Dotted lines indicate the locations of roughness measurements)………………………………………. 51 Structure of data acquisition and control system of the pin-ondisk tribometer………………………………………………….. 55 Figure 4.19 45 47 Figure 4.20 User-interface of the control software of the pin-on-disk tribometer……………………………………………………….. 56 Figure 5.1 Friction coefficients of neat PEEK, neat PTFE and PTFE/PEEK composites as a function of time………………………………... 58 Figure 5.2 Coefficients of friction of neat PEEK and PTFE/PEEK composite as a function of PTFE volume fraction……………... 59 Figure 5.3 Wear (height) loss as a function of sliding time for PTFE/PEEK composite, neat PEEK, and neat PTFE…………………………. 60 Figure 5.4 Specific wear rates of neat PEEK and PTFE/PEEK composites as a function of PTFE volume fraction…………………………. xiv 61 Figure 5.5 Relationship between specific wear rate and friction coefficient of the PTFE/PEEK composite from current experiments………. 63 Figure 5.6 Experimental results from [28] compared with power-law predictions………………………………………………………. 65 Figure 5.7 Experimental results from [19] compared with power-law predictions………………………………………………………. 66 Figure 5.8 Experimental results from [63] compared with power-law predictions………………………………………………………. 67 Figure 5.9 Specific wear rates and friction coefficients of PTFE/PEEK composite obtained from experiments with different test methods…………………………………………………………. 68 Figure 6.1 Figure 6.1 Micrographs of transfer film on the counterface (C10 sliding on 1018 carbon steel counterface)……………………… 71 Figure 6.2 A method for analyzing images of counterface micrographs to determine the transfer film covered area……………………….. 72 Figure 6.3 Optical (non-polarized) images of steel counterface after wear tests of (a) Neat PEEK polymer, (b) PTFE/PEEK composite (C20) and (c) Neat PTFE polymer……………………………… 74 Figure 6.4 Optical (polarized light) images of steel counterface after wear tests of (a) Neat PEEK polymer, (b) PTFE/PEEK composite (C20) and (c) Neat PTFE polymer…………………………........ 74 Figure 6.5 Transfer films area coverage ratio for composites with different PTFE volume fractions…………………………………………. 76 Figure 6.6 XPS scan spectra of virgin and tested steel counterface sliding over the PTFE/PEEK composite………………………………... 77 Figure 6.7 XPS C 1s spectra of virgin and tested steel counterface………... 77 Figure 7.1 PTFE/PEEK composite friction coefficients and mechanical properties……………………………………………………….. 80 Predicted μc (from Eq. 7.1) and test results on friction coefficients of PTFE/PEEK composite………………………… 81 Comparison of PTFE/PEEK composite friction coefficient predictions with experimental results…………………………... 87 Figure 7.2 Figure 7.3 xv Figure 7.4 Figure 8.1 Figure 8.2 Figure 8.3 Figure 8.4 Figure 8.5 Comparison of specific wear rate solutions with experimental results (β=4)…………………………………………………….. 90 Friction coefficients during sliding wear of neat PEEK, neat PTFE, C10 and C15 composites at 60 and 200 °C……………... 92 Wear of PEEK, PTFE, C10 and C15 composite at 60°C and 200°C…………………………………………………………… 94 (a) Steel counterface (b) sample worn surface and SEM images of (c) steel counterface and (d) worn surface of neat PEEK after friction and wear sliding test at 60°C…………………………... 96 (a) Steel counterface, (b) sample worn surface and SEM images of (c) steel counterface and (d) worn surface of neat PEEK after firciton and wear sliding test at 200°C………….……………… 96 (a) steel Steel counterface, (b) sample worn surface and SEM images of (c) steel counterface and (d) worn surface of C15 composite after friction and wear sliding tests at 60°C….…….. 97 Figure 8.6 (a) Steel counterface, (b) sample worn surface and SEM images of (c) steel counterface and (d) worn surface of C15 composite after friction and wear sliding tests at 200°C…………………… 98 Figure 8.7 Steel counterface, (a) and (d); PTFE sample worn surface, (b) and (e), and micrographs of counterface, (c) and (f). (a), (b) and (c) from tests at 60°C, and (d), (e) and (f), from 200°C………... 99 Power law relationship between friction and wear at various temperatures, (a) below Tg of PEEK matrix (152 °C), (b) above Tg of PEEK matrix……………………………………………... 102 Figure 8.8 Figure 8.9 Values of β in the power-law relationship and the storage modulus of neat PEEK………………………………………….. 104 Figure 8.10 PTFE crystallite structure (following the illustrations in [5])…... 105 Figure 8.11 Schematic PTFE sliding mechanisms (following the illustrations in [7] and [5])……………………………………… 106 Approximate relation of and Vf of PTFE in PTFE/PEEK composite……………………………………………………….. 108 Figure 8.12 xvi Figure 8.13 Comparison between friction theory (Eq. 8.4) and current experimental results…………………………………………….. 110 Figure 8.14 Comparison between friction theory (Eq. 8.4) and experimental results from [8]………………………………………………….. 110 Figure 8.15 Comparison between friction theory (Eq. 8.4) and experimental results from [3]………………………………………………….. 111 Figure 8.16 Comparison between friction theory (Eq. 8.4) and experimental results from [7]………………………………………………….. 111 Figure 8.17 Theoretical and experimental friction coefficients of neat PEEK, neat PTFE and PTFE/PEEK composites at elevated temperature……………………………………………………... 113 Specific wear rates of PTFE/PEEK composites (C10, C15, and C20) and theoretical predictions………………………………... 115 Figure 8.18 Figure 9.1 SEM image of the trailing edge of PEEK slid at 200 °C……….. 118 Figure 9.2 SEM image (back-scattering mode) of transfer films of C15 slid on steel counterface at 200°C…………………………………... 121 Figure A.1 Flow chart of the particle reinforced composite model………… 144 Figure A.2 Illustration of slicing of the cross section………………………. 145 Figure A.3 A cross section of 10% by volume particle filled composite…... 145 Figure A.4 Distribution of the area to volume fraction ratio of a 10% composite for different particle mean radii……………………... 146 xvii List of Tables Table 4.1 Mechanical properties of PEEK, PTFE and PTFE/PEEK composite materials at room temperature…...………………….. 39 Experimental matrix for room temperature friction and wear tests……………………………………………………………... 52 Table 4.3 Elevated Temperature Tribological Test Matrix……………….. 53 Table 5.1 Test conditions of PTFE/PEEK friction and wear experiments from the literature………………………………………………. 65 Table 4.2 Table 7.1 Mechanical properties and friction coefficients of neat PEEK, PTFE and PTFE/PEEK Composites at room temperature…...…. 79 Table 8.1 Results of friction and wear experiments of neat PEEK, neat PTFE and the PTFE/PEEK composites in sliding contact at different temperatures…………………………………………... 92 Experimental and theoretical predictions of friction and wear of neat PEEK and PTFE/PEEK composites, and the exponent β in power law relation……………………………………………… 101 Parameters used for Eq. 8.4 to determine composite friction coefficient………………………………………………………. 109 Parameters used for Eq. 8.5 to determine composite wear rate… 114 Table 8.2 Table 8.3 Table 8.4 xviii Chapter 1 Introduction Tribology is a discipline that studies surface interaction of contacting materials in relative motion, which includes friction, wear and lubrication. The first appearance of the term “tribology” is in the report by Jost [1] in 1966. The prefix “tribo”, which means rubbing, is derived from the Greek language. Together with the suffix “ology”, the literal translation of “tribology” would be “the science of rubbing”. From International Space Station to common household appliances, tribology plays an important role in almost all mechanical devices and structures with contacting surfaces in relative motion. Loss due to friction and wear lowers the efficiency of machinery; causes unexpected damage and repair and shortens their service life. According to the Jost Report [1], 1% of the total gross domestic product (GDP) of Britain was lost due to friction and wear of machinery in 1966; not counting the financial losses associated with machine down-time which could be even more costly than machine maintenance and parts replacement. A recent report [2] published by the United States Department of Energy on tribology opportunities for enhancing America’s energy efficiency identified 2.1% of the GDP of energy loss related to friction and wear could be saved annually. In 2017, Holmberg and Erdemir [3] concluded that about 23% of the world’s total energy loss originated from tribological contacts. Among them, 20% was used to overcome friction and 3% was used to remanufacture worn parts and spare equipment due to wear and wear-related failures. Thus, tribological considerations are of critical concerns to design and material selection for moving contacting surfaces in mechanical systems and structural components. This includes multiple factors such as temperature of contacting surfaces, their relative sliding speed, contacting loads and surface geometry. Without careful tribological considerations 1 in design and material selection, extensive maintenance may be required during service and unexpected catastrophic failure may occur. Applications of tribological principles have had a long history, though the use of the scientific term “tribology” and its study as an engineering discipline are still relatively young. In 1880 B.C., ancient Egyptian poured liquid (likely water) into the path of a moving sledge to reduce friction while transporting a heavy statue [4]. At the end of the 15th century, the beginning of European maritime exploration, Leonardo Da Vinci (14521519) firstly introduced the concept of fiction coefficient (π). Observing sliding of a wooden block on a flat surface, he concluded that the friction coefficient is the ratio of tangential (frictional) force to normal load on the wooden block and is independent of the magnitude of the normal load [5]. In 1699, two important rules of friction were proposed by Guillaume Amontons (1663-1705), who investigated the sliding of two flat surfaces against each other [6]. Similar to Da Vinci’s discovery, the first rule stated that frictional force is directly proportional to the load applied perpendicular (normal) to the sliding direction. The seconded rule, according to Amontons, is that friction coefficient is independent of the apparent contact area of the two sliding bodies. Later in 1785, French physicist Charles-Augustin de Coulomb (1736-1806) added the third rule of friction, stating that frictional force is independent of the sliding speed once motion started [7]; thus a clear distinction between kinetic and static friction is drawn. Comparing to friction, the study of wear has a shorter history. After the industrial revolution, wear issues in mass production machinery and railroads were recognized but solutions to wear problems were mostly empirical. With rapid industrial growth in the 20th century, the study of both friction and wear in metallic materials has rapidly accelerated and achieved significant 2 advancements, evidenced by the development of contact mechanics theories [8], various wear mechanisms [9, 10], semi-empirical friction [11] and wear models [12] and lubrication theories [13] for metallic materials. In the past few decades, thermoset and thermoplastic polymers are increasingly used for various engineering applications, due to their outstanding strength/density ratio, chemical resistance, and machinability. The friction and wear behavior of engineering polymers and polymer composites has also been actively investigated [14–20]. Most engineering polymers has high friction and wear which limited their application as tribological-worthy materials. To overcome the disadvantages of the polymers, particulate- and short-fiber reinforced polymer composites, with superior tribological performance than neat polymers, were introduced [21]. Adding short fibers into polymer matrix mainly improved the mechanical strength of a polymer composite. Other added inorganic and organic particles, like graphite and PTFE, lubricate the composite when sliding against another surface. The particles were considered as solid-state lubricants and substantially reduced the friction coefficient of the polymer. Reduction in friction was believed to be associated with transfer film formation on the harder surface of the counterface [22–25]. Recently reported results on modeling the friction and wear behavior of polymer composites suggested that friction reduction will ultimately lead to higher wear resistance of polymer composites [26]. Among many tribological polymers and composites, PTFE/PEEK composite is of particular interest due to its excellent tribological performance and outstanding mechanical strength and chemical resistivity [19]. Both PEEK and PTFE are semi- 3 crystalline thermoplastics. Chemical structures of neat PEEK and PTFE are shown in Figure 1.1. Figure 1.1. Chemical structures of PTFE and PEEK polymers. The PEEK polymer serves as the composite matrix due to its excellent combined mechanical properties and thermal stability. The PTFE polymer possesses outstanding lubrication properties and acts as a solid-state lubricant inside the PEEK matrix [19]. Even though the tribological performance of PTFE/PEEK polymer composite has been studied and reported by many investigators [19, 27–29], the ability to predict, either analytically or empirically, the friction and wear rate of PTFE/PEEK composite is still lacking. In view of the wide use of PTFE/PEEK composite for tribological applications, predictive models (analytical or engineering models) are needed to help understand 4 friction and wear mechanisms of the composite for accelerated development and applications of the composite for tribological operations. The lack of knowledge on the composite wear and frictional behavior introduces serious barriers in material development, manufacturing, design and consequently, prediction of their reliability in long-term applications. Issues and difficulties in determining the PTFE/PEEK composite tribological behavior and understanding its complicated mechanisms mainly result from inherently inhomogeneous microstructures of the composite, interactions among the particles, the surrounding matrix and the interface between them. Transfer film formation, chemical reactions during sliding, and lack of quantitative evaluation of the transfer films introduce additional complexities. Also, many parameters, including sliding speed, pressure, temperature, PTFE volume fraction, and transfer film morphology, have an influence on the tribological behavior of the composite and need to be considered. The objective of this study is to conduct a comprehensive investigation on the tribological behavior of PTFE/PEEK composite at both room and elevated temperatures. A combined theoretical modeling and experimental investigation is used to resolve several critical issues that have not been successfully addressed regarding friction and wear of PTFE/PEEK composite: (1) the relationship between friction and wear at room and elevated temperatures; (2) the role of transfer films on the friction coefficient of PTFE/PEEK composite; (3) the tribological behavior of PTFE/PEEK at elevated temperatures; and (4) quantitative predictive models for friction and wear at elevated temperatures. 5 A comprehensive experimental program is developed first and later supports the analytical modeling of friction and wear. Experiment efforts include: (1) pin-on-disk tribological testing of PTFE/PEEK composite at various temperatures with a special inhouse design and fabricated heating and temperature control device; and (2) new methodology for quantitative measurements of the areal coverage of transfer film. Theoretical efforts include: (1) development of the power law of friction and wear of PTFE/PEEK composite; and (2) derivation of a semi-empirical friction and wear model of the composite at room and elevated temperatures. Numerical simulation is also conducted to establish the relationship between the cross-sectional area fraction and volume fraction of randomly distributed particulate composite. Such relationship is essential to the development of friction and wear model for the PTFE/PEEK composite. In the next chapter, a review of literature is conducted on the previous work relevant to the current study, including self-lubricating polymer composites, friction and wear of PTFE/PEEK composites, experimental methods and state-of-the-art theoretical advances in polymer composite tribology. In Chapter 3, it provides a description of the objective and detailed scope of this research. Chapter 4 includes a detailed experimental program and materials information employed in this study. The results and discussions are presented Chapters 5, 6, 7, 8, and 9, addressing the aforementioned critical tribological issues of the PTFE/PEEK composite. Specifically, correlations between friction and wear of PTFE/PEEK composite are presented in Chapter 5. Also included in this chapter is the development of an empirical predictive wear model for the PTFE/PEEK composite and validation of the wear model with test results obtained in this study and from those available in the literature. In Chapter 6, details of the effect of 6 transfer films on the friction and wear behavior of polymer composite are given. An experimental method for transfer film characterization method is also described. Chapter 7 addresses the development of composite friction models that take into considerations of mechanical properties of neat polymers and the composites, and the presence of transfer films on the counterface. Also included in this chapter is verification of the models with test results. Chapter 8 presents the tribological study of PTFE/PEEK composite at elevated temperatures. Models developed in Chapters 5 and 6 are included in this chapter as references and compared with experimental data. Chapter 9 discussed microscopic friction and wear mechanisms of the PTFE/PEEK composites at elevated temperatures. Conclusions based on the experimental results acquired and the predictive models developed in the study are discussed in Chapter 10. 7 Chapter 2 Literature Review 2.1 Polymer Composites with Solid-State Lubricants Engineering semi-crystalline polymers have been used in many applications for their high strength-weight ratio, good corrosion resistivity, thermal stability and machinability [30]. For rotating and reciprocating machinery such as pumps and drilling equipment, polymers are employed commonly as sealing materials [31]. Besides the mechanical and chemical properties such as stiffness, strength, thermal and chemical stabilities, tribological properties, i.e. friction coefficient (π) and wear rate (αΊ), are also critical to polymers used in such applications [13]. Literature reported friction coefficients and wear rates of common polymers are collected and plotted in Figure 2.1 as blue circles. In Figure 2.1, friction and wear data of PPS, PES and PSU (0.2 m/s, 50N in air) are cited from [33]. Data of POM, PETP, PEI and PA-66 (0.25 m/s, 50N in air) are cited from [34]. Data of UHWMPE (50Hz, 1.1 MPa, 20°C) are cited from [35]. Data of PFA (0.0508 m/s, 6.3 MPa in air) cited from [36]. Data of PS (0.431 m/s, 50N, 25°C) are cited from [37]. Data of PET (0.025m/s, 250N, room temp.) are cited from [38]. Data of PEEKK, PEN, PI and PA-66 (1 m/s, 1 MPa, 20°C) are cited from [39]. Data of Epoxy (0.4 m/s, 3 MPa) are cited from [40]. Data of Al2O3/PTFE (0.05 m/s, 260 N, room temp.) are cited from [41]. Data of Gr/PI, SCF/PI, Gr/SCF/PI, and SiO2/Gr/SCF/PI (0.431 m/s, 200N, room temp.) are cited from [42]. Data of PTFE/PPS and Al2O3/PTFE/PPS (2m/s, 200N, 24°C) are cited from [43]. Data of TiO2/PEEK, PBI and PPP (0.1m/s, 1MPa, room temp.) are cited from [44]. A goal of material design is to get close to the lower left corner of the figure (shaded by the color of blue) where combined low wear and friction are desired. As shown in Figure 2.1, most polymers are poor 8 tribological materials. They suffer from either high wear or high friction. For example, PTFE has outstanding fiction (μ < 0.2) but high wear (αΊ >10-4 mm3/Nm). PBI, quite the opposite, suffers from the high friction (μ > 0.7) but enjoys a low wear rate (αΊ <10-6 mm3/Nm). Neither PTFE nor PBI is favorable for tribological applications. It is a common problem of polymers that they normally do not possess a combination of low friction and wear rate. Therefore, solid state lubricants are introduced to improve tribological performance [32]. Figure 2.1. Specific wear rate and friction coefficient of selected neat polymers (blue dots) and polymer composites (orange dots) sliding against a metal conterface. 9 Solid-state lubricant, normally exhibiting low friction during sliding, can be either organic or inorganic substance. It has several advantages, such as light weight, easy maintenance, low cost and contamination comparing to traditional liquid-phase lubricant in high pressure and temperatures environment [45, 46]. In addition, when using solidstate lubricant, equipment required for circulating liquid lubricant like pumps, filters and reservoirs are no longer necessary [46, 47]. Common inorganic solid-state lubricants for polymeric composites include graphite, molybdenum disulfide (MoS2) and others [48]. The structures of them are basically lattice structures with two-dimensional sheets stacking on each other and bonded by week van der Waals force as illustrated in Figures 2.2 (a) and 2.2 (b). Each layer can easily slide with respect to other layers with little resistance to the motion resulting in low friction. Typical organic solid-state lubricants are polytetrafluoroethylene (PTFE) and high-density polyethylene (HDPE) [49]. Molecular structures of all polymeric solid-state lubricants share a common feature – only small single atoms (e.g., hydrogen or fluorine) are attached to carbon backbones directly without any side chains, as shown in Figure 2.2 (c) (for PTFE). Small side atoms give minimum steric hindrance and allow polymer chains pack tightly, which lead to low surface energy and may slip easily over other chains. This feature yields low shear strength and large plastic flow during sliding and eventually achieves low friction. Unlike inorganic lubricants, yielding and large plastic flow result in material transfer from the bulk polymer to the mating surface, which leave thin-layer of transfer films [50]. The transfer films left on the counterface are proved to have significant impact on reduction of polymer fiction, wear, or both [51]. 10 As discussed previously, polymers suffer from high wear or friction for most engineering applications. So, organic and inorganic solid-state lubricants are filled into neat polymer to improve their tribological performance. Tribologically modified polymers, such as PTFE/PEEK, Gr/PI, and Al2O3/PTFE composites, exhibit excellent friction and wear properties as shown in Figure 2.1 (orange dots). Compared to friction and wear of neat polymers (blue dots), both wear and friction for the polymer composites are reduced. Although, some reduction in mechanical properties due to the incorporation of soft solid-state lubricants are observed, they are outweighed by substantial improvements on tribological performance. Figure 2.2. Lamellar lattice structures of (a) graphite (b) molybdenum disulfide and (c) polymer chain of polytetrafluoroethylene. 11 Instead of introducing solid-state lubricants into hard polymers, another method is to incorporate hard particles into soft lubricous polymers like PTFE. The friction coefficient of neat PTFE is low, but its wear is extremely high. The poor wear resistance of PTFE is attributed to large sheet-like polymer continuously transferred to the counterface, which then gets expelled from the sliding interface [51, 52]. The wear resistance of PTFE composites reinforced with hard particles or fibers, such as alumina, glass fibers, bronze or titanium oxide, has been improved by two orders of magnitude [51, 53–57]. Although wear reduction mechanisms of hard-particle reinforced composites are still unclear, several hypotheses are proposed in the literature. The first is that particles in the composites arrest propagating cracks. Blanchet and Kennedy [51] claimed that severe wear of neat PTFE was a consequence of subsurface delamination. The role of the hard particles in PTFE is to interrupt subsurface deformation and to prevent crack propagation, which otherwise produce large wear debris. The second hypothesis states when fibers and particle fillers of suitable size are introduced into soft polymers, wear is reduced because of the load-supporting action of particles. Tanaka and Kawakami [56] explained the load-supporting action of fibers and hard particles theoretically by modifying the analysis for effects of discontinues fibers in strengthening the composite. The third hypothesis on nano-sized particle filled polymers claims that nanoscale particles help improve transfer film uniformity and tenacity. Experimental results of nanoscale ZnO [58] and alumina filled PTFE [41] have been reported supporting this hypothesis. A more recent study compared friction and wear of micro- and nano- sized alumina filled PTFE [54]. It concluded that friction was less affected by the size of the alumina, but the wear resistance was significantly improved by nano-sized alumina 12 because it helped form thin uniform transfer films. Contrary to the third hypothesis, Kandanur et al. [59] reported not all nanoscale particles were effective on reducing wear. They compared various micro- and nano-scale particles reinforced PTFE and discovered that the wear reduction introduced by micro-particles was diminished when the size of particle reduced to a few tens of nanometers. To date, functionalities of hard particles and wear reduction mechanisms are still in debate. Further investigation is required for a better understanding of how particles reduce the wear of PTFE and other similar polymers. 2.2 Friction and Wear of PTFE/PEEK Composite Among many tribological polymer composites, PTFE/PEEK is commonly used in rotating and reciprocating machinery and serving as a matrix for further compounding with carbon fibers, graphite flakes and other fillers. In the literature, several papers address tribological behavior of PTFE/PEEK composite. Briscoe [60] reported a study on wear and friction of PTFE/PEEK polymer composite in pin-on-disk tests. PTFE fillers in the composite reduced friction coefficient of the composite but increased the wear rate slightly. Although this is the earliest attempt to measure friction and wear of PTFE/PEEK composite, the wear rate obtained in this study is contradictory to later studies. Lu and Friedrich [28] performed sliding tests on PTFE/PEEK composite against 100 Cr6 steel in a pin-on-disk tester and reported that both friction and wear of the composite decreased with increasing PTFE content. Hufenbach et al. [29] test PTFE/PEEK sliding against a smooth steel surface and determined the optimum fraction of PTFE for tribological applications is 7.5% (by weight). In addition to sliding PTFE/PEEK composite on a smooth surface, Bijwe and co-workers [61] study the tribological behavior of 13 PTFE/PEEK composite sliding against abrasive surfaces and find that specific wear rate of the composite increases with increasing PTFE weight fractions. They compare abrasive wear test results with mechanical properties of PTFE/PEEK composite and conclude that hardness and tensile strength are dominating wear-controlling material properties in the case of abrasive wear. Burris and Sawyer [19] study filled PEEK particles in a PTFE matrix and show that the PEEK particles decrease the wear rate and friction of the composite. The synergetic effect on friction and wear of PTFE/PEEK composite is attributed to networking and nanoscale penetration of PTFE through PEEK particles. In another study [27], they also fabricate a compositionally graded PEEK/PTFE composite with one end is PTFE-rich and test it using different methods (linear reciprocating, rotating pin-on-disk, and thrust washer). They discover the compositionally graded composite reduces both friction and wear comparing those to the bulk component without sacrificing mechanical properties. Lal et al. [62] investigate the tribological performance of PTFE/PEEK composite in harsh environments. The inclusion of PTFE in PEEK improves lubrication and wear resistance. The geometrical effect of PTFE fillers on tribological properties is studied by Vail et al. [63]. Vertically oriented PTFE fibers reduce wear by an order of magnitude comparing to PTFE powder filled one. Qu et al. [64] study the PTFE/PEEK composite with different PTFE contents and observe a synergistic effect of PTFE (up to 20% by volume) in PEEK in reducing friction and wear simultaneously. The reduction of wear and friction may result from transfer film formation on the steel counterface during sliding. Further, Qu et al. [65] also study elevated temperature friction and wear of PTFE/PEEK composite at 200°C. Tribological performance are compared below and above the PEEK glass transition temperature. At 14 200°C, plastic flow of the composite develops during sliding which leads to thick transfer films. The wear mechanism is therefore switched from adhesive wear to a plastic-flow dominated wear. 2.3 Effect of Transfer Films on Tribological Properties of PTFE/PEEK Composite Though many fundamental investigations of polymer and polymer composite tribology have been done with the aim at improving their mechanical and tribological properties, transfer films in polymer friction are critical and worthy of great attention. Polymer transfer films on mating counterface during sliding have been observed by many researchers [24, 52, 66, 67]. 2.3.1 PTFE PTFE transfer films on a glass substrate and its friction mechanisms have been studied by Makinson and Tabor [68]. They observe that at the beginning of sliding contact, the friction is high (µ>0.1). However, the friction coefficient of PTFE drops to below 0.07 after the initial period of time and find that within the initial period of sliding, large lumps and slabs of PTFE adhere to the glass substrate. After sliding for a while, thin transfer films of 0.1 – 0.4 µm thickness are observed on the counterface instead of lumpy debris. Pooley and Tabor [49] further study the transfer of PTFE to glass substrate and reveal more details of drawing of transfer films. The authors claim that drawing of transfer films of PTFE is an intra-crystalline process, and the massive fracture of drawing to be inter-crystalline. The transfer film formation process of PTFE is summarized by Biswas and shown in Figure 2.3 [69]. In the beginning of sliding, lumpy transfer is occurred due to high adhesion between the counterface and the polymer. Thin films are 15 then formed due to drawing of the lumps and slabs. With the presence of thin transfer film, interaction between PTFE and the counterface surface becomes the interaction between the bulk PTFE and transferred films. The latter interaction keeps the thin oriented films to be drawn continuously out from the PTFE by a shear. The plastic deformation and drawing of PTFE require lower shear forces and energy than those associated with bulk failure, which results in a low steady-state friction of PTFE. Figure 2.3. Schematic model of transfer film development of FTFE [69]. 16 2.3.2 PEEK Unlike PTFE, thin and continuous transfer films are not observed for neat PEEK. Voort and Bahadur [70] study the sliding behavior of neat PEEK under test conditions of 19.6 N normal load, 1.0 m/s sliding speed and 0.11 μm center-line surface roughness. Figure 2.4 shows steel counterfaces (Ra = 0.11 μm) during different sliding traverses of neat PEEK. Up to 100 cycles, only small amount of PEEK (appears dark in color in the image) deposits in the valleys of steel asperities. Figure 2.4. SEM pictures of steel counterface after (a) 1 cycle; (b) 10 cycles; (c) 100 cycles; (d) 10,000 cycles; (e) 70,000 cycles; and (f) 141,000 cycles sliding of PEEK. 17 After 10,000 cycles, patches and clumps of transferred PEEK start to show. With more cycles, the same process occurs repeatedly, covering more groves over a larger area. However, most of the counterface is still not covered with PEEK patches of any significant thickness after 141,000 cycles. The lack of a continuous transfer film of PEEK indicates that adhesion between the polymer transfer film and the steel surface is weak. 2.3.3 PTFE/PEEK Composite Transfer films of PTFE/PEEK composite are discussed in two cases, depending on the amount of PTFE in the composite. Adding a low volume fraction of PTFE into PEEK matrix and introducing PEEK particles into PTFE matrix result in different microstructures. The continuous phase of these two composites are PEEK and PTFE, respectively. Burris and Sawyer [19] investigate both cases and discover for the case of PTFE filled PEEK, the PTFE is drawn out of the matrix to form transfer films on the counterface lubricating the contact surface. Below a specific amount of weight fraction, the spacing between the PTFE particules become large such that transfer films cannot completely cover the steel counterface to lubricate the PEEK. In the case of introducing PEEK particles into PTFE matrix, with an increasing amount of PEEK, the PTFE friction coefficient is reduced due to the shearing of low strength running films over an increasingly stiff material with less real area in contact. Qu et al. [26] also point out that for a PTFE/PEEK composite with less than 20% PTFE by volume, the area covered by transfer films increase with PTFE content. The increase in transfer film area is believed to result in a decrease in friction coefficient. 18 2.4 Experimental Methods for Polymer Tribological Study The friction force and volumetric wear loss of a material during sliding contact may be measured by a tribometer. The experimental apparatus normally holds a test specimen mating with a stationary substrate or vice versa. The fiction force between the sliding pair is measured by a load cell and the materials loss is determined by a linear variable differential transducer (LVDT). Commercially available tribometers generally allow control of experimental parameters, such as normal load, sliding velocity, specimen geometry, temperature and humidity. Bayer [71] and Benzing et al. [72] review various friction and wear testers that have been employed for tribological evaluation. Experimental systems commonly used for polymer tribological tests are shown in Figure 2.5. Although these test systems are most commonly used, actual tribometers used by different researchers may be custom-built machines. Instead of preforming tribological tests with a tribometer, some tribological experiments may be performed directly on the full-size machines under real application conditions. The following subsections discuss the features of each selected tribo-test system. 2.4.1 Block-on-Ring Tribometer In the block-on-ring apparatus shown in Figure 2.5 (a), the sample block is held stationary and the ring rotates. A load is applied perpendicular to the test specimen along axis of rotating ring. The sample block may be cylindrical or cubical. One end contacting the ring has an arc of the same radius of the ring. This method is standardized by ASTM G137 – 97 [73]. 19 Figure 2.5. Common tribological test systems: (a) block-on-ring tribometer, (b) pin-ondisk tribometer, (c) linear reciprocating tribometer and (d) thrust washer tribometer. 2.4.2 Pin-on-Disk Tribometer The pin-on-disk tribometer, as shown in Figure 2.5(b), may be the most popular test system for polymer tribological experiments. In this apparatus, either the pin is held stationary with a rotating disk or vice versa. The sample pin can be a ball or a cylinder with a flat end. The disk rotates at a constant speed during sliding contact with a normal load applied on top of the specimen pin. The procedure for using the test method is standardized by ASTM G99 – 17 [74]. 20 2.4.3 Linear-Reciprocating Tribometer In a linear reciprocating test apparatus, as shown in Figure 2.5(c), the specimen pin is stationary and the flat substrate moves in reciprocating motion. The geometry of the test specimen can be a ball, cylinder, cube or even parallelepiped. The substrate can also oscillate with a small amplitude for fretting wear experiments. This test method is standardized by ASTM G133 – 05 [75]. 2.4.4 Thrust-Washer Tribometer In the thrust-washer tester, as shown in Figure 2.5(d), the stationary flat counterface is a washer/disk. The sample thrust rotates or oscillates on the washer surface. The load is applied parallel to the axis of rotation. The test method is standardized by ASTM D3702 – 94 [76]. 2.5 Friction and Wear Theories of Polymer Composite 2.5.1 Friction When a particulate-reinforced composite slides against a steel counterface, the total normal load πΏ is taken by the particles πΏπ and by the matrix πΏπ along the contacting surface, πΏ = πΏπ + πΏπ . (2.1) On any cross section of the composite, particles are randomly distributed across the surface. The relationship between the area fraction (π΄π ) and the volume fraction (ππ ) of the particle in the composite is approximately equal, π΄π ≅ ππ . 21 (2.2) Detailed proof and numerical simulation of the relation between cross-sectional areal and volumetric fraction of a randomly distributed particulate-reinforced composite is in Appendix A. Assume that the contact shear stress (π) caused by friction across the crosssection area is uniformly distributed, i.e., π≅ πΉ πΉπ πΉπ ≅ ≅ , π΄ π΄π π΄π (2.3) where πΉ is the friction force and π΄ is the contact area. Subscripts, π and π , indicate particles and the matrix of the composite, respectively. Note that π΄π ππ π΄ and π΄π (1 − ππ )π΄. Since the friction coefficient is defined as, ππ = πΉ/πΏ, Eq. 2.1 may be written as πΉπ πΉπ πΉ = + , ππ ππ ππ (2.4) where ππ = πΉπ /πΏπ and ππ = πΉπ /πΏπ . Substituting Eq. 2.3 into Eq. 2.4, ππ 1 − ππ 1 = + . ππ ππ ππ (2.5) Eq. 2.5 suggests an inverse rule-of-mixtures type friction equation for the composite. Similarly, if uniform pressure along the contact surface is assumed, ,a linear rule-of-mixtures type friction equation is obtained, ππ = ππ ππ + (1 − ππ )ππ . 22 (2.6) 2.5.2 Transfer Films The solid film-lubrication theory by Rabinowicz [13] introduces the effect of transfer films on sliding friction. In a non-lubricated sliding contact i.e., dry sliding of polymer without a lubricating film as shown in Figure 2.6, the friction coefficient ππ may be determined by the ratio of shear yield stress and hardness of the softer material. In the case of a polymer-metal contact sliding, the polymer is the soft material, thus ππ = ππ , π»π (2.7) where ππ is the shear yield stress and π»π is the hardness of the polymer. When a soft lubricating film is developed between the polymer and the steel counterface, as illustrated in Fig. 2.6 (b), sliding motion takes place either in the film or along the film-solid interface. In either case, the friction coefficient of the lubricating film can be shown to be ππ = ππ , π»π (2.8) where ππ and ππ are, respectively, the friction coefficient and shear stress of the lubricant film. Taking the ratio of ππ to ππ , we have ππ ππ ππ = ππ ππ = ⋅π . π π ππ ππ π 23 (2.9) Figure 2.6. Solid film lubrication at a polymer-steel contact junction. If the lubricating film is assumed to have the same tribological and yield properties of the composite, then the composite friction coefficient may be expressed in terms of friction coefficient of the neat polymer matrix. Thus Eq. 2.9 may be rewritten as ππ = ππ ⋅π . ππ π (2.10) Eq. 2.10 gives a simple relation between friction coefficient and shear yield stresses of the composite and the matrix polymer. 2.5.3 Wear Polymer-matrix composites for tribological applications normally contain short reinforcing fibers and particles and solid-state lubricants. To better understand the relation between wear and the composition of a polymer composite, phenomenological wear theories have been developed [77, 78]. Assuming fillers are distributed uniformly and non-interactive in the matrix, the rule-of-mixtures may be applied to approximate the wear rate of the composite. Since a unit volume of a polymer composite is represented by the sum of volume fraction of the fillers and the matrix 24 , the total normal load is the sum of the loads on the fillers and the matrix . The normal load on the matrix and on the fillers can be approximated as πΏπ = ππ ⋅ πΏ and πΏπ = (1 − ππ ) ⋅ πΏ (2.11) if contact pressure is assumed to be uniform and the (cross-sectional) area fraction is assumed equal to the composite volume fraction. Therefore, the specific wear rate of a composite may be obtained in terms of the volume fractions of individual phase materials. Khruschov [77] introduces a linear rule-of-mixtures wear model for composite as π€Μπ = ππ π€πΜ + (1 − ππ )π€π Μ , where , , and (2.12) are specific wear rates of the composite, the reinforcing particle and the matrix, respectively. The model successfully predicted the wear of a brass/lead composite [77]. However, it failed to predict other composite, such as tungsten carbide/cobalt (a hard, brittle and porous composite). Based on Khruschov’s wear equation, the inverse rule of mixture model of abrasive wear is proposed by Simm and Freti [78], ππ 1 − ππ 1 = + . π€πΜ π€πΜ π€π Μ (2.13) The two linear rule-of-mixtures and inverse rule-of-mixtures wear models, Eqs 2.11 and 2.12, are derived for abrasive wear of metal composites. Transfer films are not considered. 25 2.5.4 Flash Temperature When two surfaces slide against each other, frictional heating raises the temperature of the interface above that of the environment. This temperature rise is termed the flash temperature raise. The maximum temperature rise π©, at the counterface during sliding contact may be estimated by Θ= 2 ππ 1 , √π πΎ √ππ (2.14) where q is heat generated, assuming all the heat comes from frictional work; a is the characteristic contact length, and K, thermal conductivity. Pe is the Peclet number, defined as ππ = ππ£ , 2π (2.15) with κ being thermal diffusivity. The heat source, q in general is the result of friction during sliding and is dependent on the applied normal pressure P, friction coefficient µ, and the sliding velocity v. Therefore, the value of q may be determined by π=πππ£ (2.16) Thus, in the subsequent evaluation and modeling of the friction coefficients, temperature effects need to be considered. Friction, wear and other relevant material properties should be evaluated at the temperature associated with the flash temperature rise. To avoid issues related with flash temperature, temperature at the interface between test sample and counterface is controlled for all high temperature tests. 26 Chapter 3 Objectives and Scope of Research Friction and wear behavior of polymer composites are complicated system responses. Despite the widely use of PTFE/PEEK composites in tribological applications, friction and wear behavior and their associated mechanisms and predictive theories are not fully studied. Fundamental knowledge of tribological properties of PTFE/PEEK composite is essential for future material development and applications. For elevated temperature applications, thermal properties are critical. The objective of this study is to understand the nature of the sliding friction and wear behavior of PTFE/PEEK composite at both room and elevated temperatures. The scope of the research includes the following items: 1. Develop a tribological testing system capable to evaluate friction and wear properties of the PEEK polymer and the PEEK-based polymer composite at elevated temperature. 2. Evaluate fundamental mechanical and thermal properties of neat PEEK, neat PTFE and PTFE/PEEK composite. 3. Room temperature tribological experiments on neat PEEK, neat PTFE and PTFE/PEEK composite with various PTFE volume fractions. 4. Elevated temperature tribological experiments on neat PEEK, neat PTFE and PTFE/PEEK composite at various temperatures. 5. Develop a quantitative method for transfer film evaluation. 6. Establish proper relationships between friction and wear of PEEK polymer and PTFE/PEEK composite. 27 7. Develop friction and wear theories for PTFE/PEEK composite at room temperature. 8. Develop friction and wear theory for PTFE/PEEK composite at elevated temperature. 28 Chapter 4 Materials System and Experimental Program 4.1 Material System 4.1.1 Constituent Materials The neat PEEK, neat PTFE and PTFE/PEEK composites investigated in this study were manufactured in the same manner as samples in previous works [26, 64, 65]. All samples were prepared by compression molding followed by sintering. Molded composite samples contained 5%, 10%, 15%, and 20% (by volume) PTFE in the PEEK matrix. The PTFE/PEEK composite were denoted, according to their PTFE volume fractions, as C05, C10, C15, C20, respectively. The neat PEEK and PTFE polymers studied were also made by compression molding from their powders. The PEEK powders were Victrex 450PF and the PTFE powders were Dakin M-15X. The actual powder size was reported in a previous study [79] to be tens of microns. 4.1.2 PTFE/PEEK Composite Microstructure Polished PTFE/PEEK samples were coated with conductive Au for scanning electron microscope (SEM) examinations. Typical images of polished surfaces of the composites were obtained by SEM secondary electron beam scanning and shown in Figures 4.1 (a)-(f). The PTFE particulates appeared in light color with irregular shapes and a major axis up to 100 µm. Phase segregation was observed and revealed the immiscible nature of the PTFE and PEEK polymers. 29 Figure 4.1. SEM micrographs of (a) neat PEEK, (b) C05, (c) C10, (d) C15, and (e) C20 composites, and (f) neat PTFE. 4.1.3 Polymer and Composite Morphology The degrees of crystallinity of neat PEEK, neat PTFE and the composites were determined by X-ray diffraction (XRD). The X-ray diffraction was performed with a Rigaku SmartLAB X-ray diffractometer, using a Cu target producing Kα X-rays (λ=1.54 Å) scanning from 10 – 30° 2θ at a rate of 5 degrees per minute. To perform the morphological study, specimens were cut to a plaque (20 mm × 20 mm × 5mm) from the bulk material and polished by 400, 600, and 1200 girt sand papers. XRD patterns of each sample, as shown in Figure 4.2, resulted from crystalline peaks of the crystalline plane superimposed on broad amorphous halos. By drawing a linear based line from 2θ 10° to 30°, XRD patterns were fitted by a Pseudo-Voigt function which was a convolution of 30 Gaussian and Lorentzian functions. The Gaussian πΊ(2π) and Lorentzian πΏ(2π) functions were expressed by (2π − 2π0 )2 πΊ(2π) = πΌπππ₯ exp [−π ] π½2 (4.1) 1 1 2Π πΏ(2π) = , π 1 2 2 (2 Π) + (2π − 2π0 ) (4.2) and where is the maximum intensity; 2π0 , the position of the peak maximum, and π½ is the integral breadth related to the full width at half maximum (FWHM), Π, by Π . The C10 composite crystalline peaks and amorphous halos fitted and separated by this method is shown in Figure 4.3. T=23 °C Figure 4.2. XRD patterns of neat PEEK, neat PTFE and PTFE/PEEK composites. 31 350000 T=23 °C Experimental Results Pseudo Voigt Fit 300000 PEEK amorphous PEEK Crystalline 250000 Intensity (ct.) PTFE Amorphous PTFE Crystalline 200000 150000 100000 50000 0 10 15 20 2θ 25 30 Figure 4.3. Crystalline and amorphous peaks of the X-ray diffraction pattern of C10. The degrees of crystallinity of neat PEEK, neat PTFE and the individual PEEK and PTFE phases in the composites were determined by integrating crystalline and amorphous peaks from X-ray diffraction traces using Ruland’s method [80] with correction factors. The degrees of crystallinity, Xc was obtained by taking the ratio of crystalline peaks areas to the total area under the curve 2π ππ = π ∫2π 2 πΌπ (2π) π2π 1 2π π ∫2π 2 πΌπ (2π) π2π 1 2π + ∫2π 2 πΌπ (2π) π2π 1 32 ⋅ 100% , (4.3) where I is intensity of X-ray diffraction. Subscripts, c and a, indicate the crystalline and the amorphous phases. The ζ is the Ryland’s correction factor which corrects the effects of polarization, diffraction angle, and temperature. The value of ζ used in the calculation was 1.0 for PEEK [81] and 1.8 for PTFE [82]. Degrees of crystallinity of PEEK and PTFE constituent in the composite are shown in Figure 4.4. Dotted lines indicate the degrees of crystallinity of neat PEEK (red) and neat PTFE (blue). The crystallinity of PEEK remained almost the same in the composite as that in the neat resin. The crystallinity of PTFE phase in the composite was about 2% higher than that in the neat PTFE. The 2% difference in crystallinity was not significant enough to cause any mechanical and thermal property differences. The degrees of crystallinity is mainly affected by the cooling rate after sintering [83]. This is associated with the processing conditions. Slow cooling in the processing normally leads to higher degrees of crystallinity. On the other hand, quenching or rapid cooling result in lower degrees of crystallinity for polymeric materials. Since all samples with different compositions were processed in the same manner and the degrees of crystallinity of each individual constituent phase was almost the constant, the effect of crystallinity was not considered in the subsequent mechanical and tribological studies. Several factors may affect the crystal size of a polymer, such as number of nucleation sites and presence of nucleating agents [83]. 33 T=23 °C Figure 4.4. Crystallinity of neat PEEK, neat PTFE, and individual PEEK and PTFE phases in the PTFE/PEEK composite. 4.2 Thermal and Mechanical Properties 4.2.1 Thermal Properties Thermal properties, such as glass transitions and melting points, of neat PEEK, neat PTFE, PTFE/PEEK composites were obtained from differential scanning calorimetry (DSC). In Figure 4.5, DSC traces of neat PEEK and PTFE resins and composite samples are shown. At 152 °C, a glass transition was observed for neat PEEK and the PEEK matrix in the composite. The PEEK matrix and the PTFE particulates in the composite exhibited same melting points as their neat resins at 340 and 332 °C, respectively. The distinct melting peaks for the PEEK and PTFE phases in their composite confirmed phase segregation in the composite. 34 EXO 0 -0.2 Heat Flow (a.u.) -0.4 -0.6 -0.8 -1 -1.2 -1.4 -1.6 PEEK C05 C10 C15 C20 PTFE PEEK Tg=152°C 100 150 PTFE melting 332°C -1.8 50 200 T(°C) 250 300 PEEK melting 340°C 350 Figure 4.5. DSC traces of neat PEEK, neat PTFE and PTFE/PEEK composites. 4.2.2 High-temperature Mechanical Properties 4.2.2 (a) Hardness Plastic yielding and failure of neat PEEK, PTFE and PTFE/PEEK composites were reported in [79]. Hardness was measured with a Rockwell hardness tester (Wilson Rockwell 2000). The procedure of hardness measurements followed the ASTM D785 Standard on Hardness of Plastics and Electrical Insulating Materials [84]. Rockwell R and M scales were applied to PTFE and other samples, respectively. A conversion from Rockwell hardness to Brinell hardness was performed by the following equation, π΅π»π = πΉπ , 0.004ππ ⋅ (130 − π π») 35 (4.4) where BHN is the Brinell hardness number; RH, Rockwell hardness; r, the indenter radius, and πΉπ , the applied force. The Rockwell hardness is determined from the depth of penetration of a ball indenter under load. Brinell hardness is also determined using a ball indenter. Unlike Rockwell hardness, Brinell hardness is determined using the area of indentation made by the ball indenter and is expressed in megapascals (MPa). Though both Rockwell and Brinell hardness are determined using ball indenters, the quantities used to characterize material hardness are different but related to each other as shown in Eq. 4.4. Selected room temperature mechanical properties of PEEK, PTFE and PTFE/PEEK composites are tabulated in Table 4.1. The measured Brinell hardness of neat PEEK, PTFE, and PTFE/PEEK composites are shown in Figure 4.6. 1000 T=23 °C 900 Brinell Hardness 800 700 H (MPa) 600 500 400 300 200 PTFE 54.9 MPa 100 0 0 0.05 0.1 0.15 Vf (PTFE) 0.2 0.25 Figure 4.6. Hardness of neat PEEK, neat PTFE and PTFE/PEEK composites at room temperature. 36 4.2.2 (b) High-Temperature Dynamic Mechanical Properties Dynamic mechanical analysis (DMA) of the aforementioned materials were performed with a TA RSA-III Thermal analyzer. Sample dimensions were 24 mm in length, 12 mm in width, and 1.5 mm in thickness. Test conditions followed ASTM D7028-07 Standard [85], which specified 1 Hz frequency and 5 ºC/min heating rate. Three-point-bending fixtures were used for the DMA tests with a fixed strain of 0.2%. Specimen were heated from room temperature to 280°C. Storage modulus, πΈ′, loss modulus, πΈ", and loss tangent, tan πΏ, were plotted as a function of temperature and shown in Figures 4.7, 4.8 and 4.9, respectively. 5×109 4×109 PEEK C05 3×109 E' (Pa) C10 C15 2×109 C20 PTFE 109 0 0 50 100 150 200 250 T (°C) Figure 4.7. Storage modulus of neat PEEK, neat PTFE and PTFE/PEEK composites as a function of temperature. 37 4×108 PEEK C05 C10 3×108 C15 E" (Pa) C20 PTFE 2×108 1×108 0 0 50 100 150 200 250 T (°C) Figure 4.8. Loss modulus of neat PEEK, neat PTFE and PTFE/PEEK composites as a function of temperature. 0.2 0.18 0.16 0.14 Tan δ 0.12 0.1 0.08 PEEK C05 C10 C15 C20 PTFE 0.06 0.04 0.02 0 0 50 100 150 200 250 T (°C) Figure 4.9. Loss tangent of neat PEEK, neat PTFE and PTFE/PEEK composites as a function of temperature. 38 39 112 91 61 49 43 12 PEEK C05 C10 C15 C20 PTFE Tension 20 80 85 98 128 143 Compression 9 49 50 54 59 76 Shear Yield Stress (MPa) [79] -- 48 50 65 93 112 Tension -- 89 102 115 128 143 Compression -- 61 62 66 66 80 Shear Failure Strength (MPa) [79] HRR HRM Scale 18.1 75.3 86 92.9 98.2 105.4 Mean 0.9 0.9 0.4 0.5 0.3 0.1 Dev. St. Rockwell Hardness 54.9 404.3 502.4 596.1 695.1 899.2 (MPa) Mean 2.7 5 2.2 3.2 1.8 0.9 Dev. St. Brinell Hardness Table 4.1. Mechanical properties of PEEK, PTFE and PTFE/PEEK composite materials at room temperature. 4.2.2 (c) High-Temperature Stiffness and Yield Stress Room temperature mechanical properties (yield stress and failure strength) of neat PEEK, neat PTFE and PTFE/PEEK composite listed in Table 4.1 are cited from [79]. Tubular sample geometry and strain gauge were used in the cited study. Elevated temperature compression tests on neat PEEK, neat PTFE and PTFE/PEEK composites were performed with an Instron ElectroPuls E10000 Linear-Torsion electrical tester. Dimensions of the test block were 6.35 × 6.35 × 12.7 mm (¼ × ¼ × ½ in.). The strain rate was 0.1%/s. Stress-strain curves of neat PEEK, C05, C10, C15, C20 and neat PTFE were shown in Figure 4.10. Elastic modulus and compressive yield stress of each specimen were obtained for all temperatures up to 200 °C. Elastic moduli and compressive yield stresses of neat PEEK, neat PTFE and the composites are plotted in Figures 4.11 and 4.12 as a function of temperature. A 2% offset yield stress was used in this study to illustrate the rapid drop in yield stress with increasing temperature. Modulus of all samples decreased with increasing temperature. At the glass transition temperature of PEEK (152 °C), a sudden drop was observed for neat PEEK and PEEK-based composites. Compressive yield stress showed a monotonic decrease with increasing temperature for all neat and composite samples. 40 -160 (a) PEEK -140 25 °C 60 °C -120 100 °C 125 °C 150 °C -80 175 °C -60 200 °C -40 225 °C -20 0 0 -0.1 -0.2 -0.3 ε -160 (b) C05 -140 25 °C 60 °C -120 100 °C 125 °C -100 σ (MPa) σ (MPa) -100 150 °C 175 °C -80 200 °C -60 -40 -20 0 0 -0.1 ε 41 -0.2 -0.3 -160 (c) C10 -140 25 °C -120 60 °C 100 °C σ (MPa) -100 125 °C 150 °C -80 175 °C -60 -40 200 °C -20 0 0 -0.1 -0.2 ε -0.3 -160 (d) C15 -140 25 °C -120 60 °C -100 125 °C 100 °C σ (MPa) 150 °C -80 175 °C -60 200 °C -40 -20 0 0 -0.05 -0.1 -0.15 ε 42 -0.2 -0.25 -0.3 -160 (e) C20 -140 -120 25 °C σ (MPa) -100 60 °C -80 100 °C 125 °C -60 175 °C 150 °C -40 200 °C -20 0 0 -0.05 -0.1 -0.15 ε -0.2 -0.25 -0.3 -50 (f) PTFE -45 -40 25 °C -35 σ (MPa) -30 -25 60 °C -20 100 °C 125 °C 150 °C -15 -10 175 °C 200 °C -5 0 0 -0.1 ε -0.2 -0.3 Figure 4.10. Stress-strain curves of elevated temperature compression tests of (a) PEEK, (b) C05, (c) C10, (d) C15, (e) C20 and (f) PTFE. 43 4500 PEEK 4000 C05 C10 EC (MPa) 3500 C15 3000 C20 2500 PTFE 2000 1500 1000 500 0 0 50 100 150 T (°C) 200 250 Figure 4.11. Elastic moduli of neat and composite samples as a function of temperature. 140 PEEK C05 120 C10 C15 100 σyc (MPa) C20 80 PTFE 60 40 20 0 0 50 100 T (°C) 150 200 250 Figure 4.12. Compressive yield stresses of neat and composite samples as a function of temperature. 44 4.3 Experimental Facilities 4.3.1 High-Temperature Pin-on-Disk Tribometer To conduct research on elevated temperature polymer composite tribology, and overcome limitations of the existing commercially available tribometers, a new hightemperature tribotester with precise temperature control was designed and built. Figure 4.13. Schematic of pin-on-disk tribometer. The tribotester was constructed with the frame of a Lewis Research LRI-a tribometer (Figure 4.13). The spindle, specimen fixtures, servo motor controller and data acquisition devices were totally replaced. With completely re-programed computer control software, 45 the newly developed tester was capable to operate at high velocity and load for a continuous long experiment. The system was driven by a servo motor with a build-in encoder which monitored and controlled rotating speed. The motor provided a speed of maximum 3500 rpm. The rotational motion was transmitted to the spindle by pulleys and a timing belt with a drive ratio of 1:0.44. A schematic of the driving train is shown in Figure 4.14. Figure 4.14. Drive train of the pin-on-disk tribometer. Friction force was measured by a load cell with a maximum capacity of 111.2N (25 lbf). The tangential (friction) force of the polymer specimen during sliding was 46 transmitted through an arm to the load cell as shown in Figure 4.15. The transmission ratio was 10/0.475. The wear loss of the sample was measured by an LVDT mounted at the end of the load beam. The LVDT converted a measured displacement into analog signals with a sensitivity of 0.394 VDC/mm. The volumetric wear loss was calculated by the specimen height loss since the cross-sectional area remained constant. Figure. 4.15. Assembly of load arm, load cell, and LVDT. 47 4.3.2 High Temperature Stage and Temperature Controller Several common methods for evaluating tribological properties of polymers and polymer composite were described previously in Section 2.4. Friction and wear rates measured by these methods show dependence on flash temperature rise at the contact surface introduced by frictional heating [86]. Tests with high load and/or speed resulted in large plastic flow or even melting of specimens due to excessive heat built up at the sliding surface [87]. In this study, a precision-test method has been developed to conduct friction and wear experiments with a controlled sliding contact surface temperature. By controlling contact-surface temperature, failure modes of the test material were ensured due to wear at a specific temperature. To accurately control of the sliding contact surface temperature, a heating stage and a control device for the pin-on-disk tribotester were designed and constructed. In Figure 4.16, schematics are shown for the temperature control assembly for a hightemperature tribology test. The fixtures consisted of a stainless-steel base for mounting the counterface disk. The base was attached to a copper cooling fin. The copper fins had different sizes to control heat removal. For temperature below approximate 80 °C, the temperature control was achieved by controlled removal of heat generated at the interface of the test sample and the counter surface. Air at room temperature (approximately 23°C) supplied by a cooling fan was passed through the cooper fin to remove heat generated at the sliding surface. 48 Figure 4.16. Temperature control stage for high-temperature pin-on-disk tribometer. In elevated temperature experiments, especially at the temperature above Tg, an augmented controlled heating is required and a heating stage was introduced (Figure 4.16). Three compartments are designed in the heating stage for cartridge heaters of choice. The sliding contact surface temperature was controlled by a temperature controller as shown in Figure 4.17. A PID (proportional–integral–derivative) algorithm was implemented to prevent temperature overshooting. In addition to the precise control of the temperature, the thermal system can record temperature profiles up to 4 thermocouple measurements. It equipped with a touch screen and a graphical user interface for easy use. 49 Figure 4.17. Multi-channel temperature controller and recorder. 4.4 Experimental Program 4.4.1 Sample Preparation Neat PEEK, neat PTFE and PTFE/PEEK composite samples were compressionmolded and followed by sintering. Specimens were cut from the bulk materials. Sample pins were machined to a 6.35 × 6.35 × 9 mm cuboid. The steel counterface was made of AISI 1018 carbon steel. Prior to each test, the sliding surface was grinded and polished by hand, using a 320-grit silicon carbide sand paper on a rotating lapping machine with a stream of tap water to remove any pre-existing machining marks. After all marks were removed, the sample surface was polished by 400 grit and a600 grit sand papers. During the surface polishing, the sample was turned at various angles to remove randomly oriented polishing lines and their effect on the sliding friction and wear tests. After 50 polishing, the surface was rinsed and wiped with alcohol. The center line roughness of the counterface was measured with a Mitutoyo SJ-210 stylus profilometer at 5 locations as illustrated in Figure 4.18. The center-line roughness of each disk was determined as the average of the 5 values measured at the indicated locations. This counterface preparation procedure resulted in a roughness (Ra) of 0.1 µm ±0.02 µm. The polished counterface provided a uniform background for subsequent studying of transfer films. Figure 4.18. A steel counterface. (Dotted lines indicate the locations of roughness measurements). 4.4.2 Friction and Wear Test Matrix A comprehensive test program on friction and wear of neat PEEK, neat PTFE and PTFE/PEEK composite was developed. Three samples were tested for each composition. A detailed test matrix is shown in Table 4.2. 51 Table 4.2. Experimental matrix for room temperature friction and wear tests. Test No. Number of Test Samples 1 3 2 3 3 3 4 3 5 3 6 3 Material Counterface Material Normal Pressure (MPa) Sliding Speed (m/s) PEEK 5% PTFE/PEEK 10% PTFE/PEEK 15% PTFE/PEEK 20% PTFE/PEEK PTFE 1018 Steel 0.5 1 (800 RPM) Normal pressure and sliding speed were chosen based on consideration of real conditions of material application (e.g., a compressor). Note that within a certain range of PV value (the product of pressure and sliding velocity), friction and wear do not change significantly. However, if the PV value is too high, different tribological behavior may be expected due to seizure and material melt down. Neat PEEK, neat PTFE and several PTFE/PEEK composites were selected for elevated temperature tribological experiments. Detailed experimental matrix is shown below in Table 4.3. The test temperatures selected in the study ranged from room temperature up to 200 °C. The experiments were expected to reveal differences in tribological behavior of neat resin and composite samples below and above the PEEK glass transition temperature (152 °C). 52 Table 4.3. Elevated Temperature Tribological Test Matrix. Test Temperature No. (°C) 1 60 2 100 3 125 4 150 5 175 6 200 Sample Material Counterface Material Normal Pressure (MPa) Sliding Speed (m/s) PEEK C10 C15 C20 PTFE PEEK C10 C15 C20 PTFE PEEK C10 C15 PTFE PEEK C10 C15 C20 PTFE PEEK C10 C15 PTFE PEEK C10 C15 C20 PTFE 1018 Steel 0.5 1 53 4.4.3 Experimental Procedure Test sample pins prepared previously were mounted on a sample holder in the pin-on-disk tribometer, with the sliding surface pre-conditioned at 0.5 MPa normal pressure under 0.1 m/s sliding against a 600-grit sand paper for 1 minute. Machining marks were removed from the sample surface by this procedure. The sample surface and steel counterface were then wiped with alcohol. Compressed air was used to evaporate alcohol residuals and remove any other solid contaminations from the sliding interface. Friction and wear test were conducted at a specified contact surface temperature in an ambient laboratory condition with ~50% relative humidity. Tests were continued for 48 hours without any interruption. The height loss of the sample, the friction force and the conterface temperature were measured with a LVDT, a load cell, and a thermocouple, respectively. 4.4.4 Data acquisition and analysis Friction, wear and temperature data were collected periodically by a personal DAQ 55 data acquisition device during the test. Data were transmitted and stored at a rate of 0.05 Hz, which yielded a 200-second logging frequency. To control and synchronize the motor speed and the temperature measurements with the data collection, a computerbased control software was developed. Figure 4.19 illustrated the test system and its control mechanisms. 54 Figure 4.19. Structure of data acquisition and control system of the pin-on-disk tribometer. The data acquisition and analysis were developed with a Microsoft Windows operating system to accommodate all aforementioned functionalities with a graphical user interface. It has a simple interaction logic and a dedicated window to print the testing status. A screen shot of the interface of the program is shown in Figure 4.20. Collected data during the test were written into a file for subsequent analysis. 55 Figure 4.20. User-interface of the control software of the pin-on-disk tribometer. The raw data obtained from a test included measured frictional force, height loss and temperature. Normal load and sliding speed were controlled and hold constant throughout the entire experiment. For each measurement, the friction coefficient was calculated by μ(π‘) = πΉπ (π‘) , πΉπ (π‘) where πΉπ (π‘) and πΉπ (π‘) were the frictional force and the normal load at time, π‘, respectively in the experiment. 56 (4.5) The height of the specimen, β, was measured and recorded periodically. The height loss rate was expressed as βΜ = πβ . ππ‘ (4.6) The specific wear rate, π€Μ , was defined as the ratio of the wear volume loss, π, and the product of the normal load and the sliding distance, π, as π€Μ = π πΉπ ⋅ π (4.7) and commonly expressed in units of ππ3 /π ⋅ π . Note that various methods are available to measure wear loss volume for determining the specific wear rate. The measurement could be direct or indirect. An indirect way is to weigh the mass of the sliding specimen and convert the mass loss to volume by dividing the density. This method usually interrupts the sliding test and a uniform density of the specimen is assumed. In the present study, a direct measurement was employed to avoid experiment interruptions and density uncertainties. Sliding samples were machined to a cuboid with constant cross-sectional area. Therefore, the wear volume loss simply became the product of the height loss and cross section area. Here, the specific wear was determined by π€Μ = βΜ ππ£ where π is the normal pressure and π£ is the sliding velocity. 57 (4.8) Chapter 5 Relationship between Friction and Wear of PTFE/PEEK Composite 5.1 Experimental Results of Friction The coefficients of friction of neat polymers and PTFE/PEEK polymer composites with different PTFE fractions were obtained from the experiments and are shown in Figure 5.1. All experiments lasted 48 hours except that on PTFE due to its high wear rate. (The test duration for PTFE was 10 hours.) The first 24 hours (4 hours for PTFE) of a friction test was considered as a running-in period. During the subsequent 24 hours of tests, friction coefficients were found to fluctuate around a constant value. Friction in this period was considered in a steady state and its associated friction coefficient was determined. 0.6 Ambient Temperature = 23 °C 0.5 PEEK 0.4 µ C05 0.3 C10 C15 0.2 C20 PTFE 0.1 0 0 10 20 t (hour) 30 40 Figure 5.1. Friction coefficients of neat PEEK, neat PTFE and PTFE/PEEK composites as a function of time. 58 Friction coefficients of neat PEEK and its composites determined in room temperature experiments are given in Figure 5.2 as a function of PTFE volume fraction. The friction coefficient decreased from 0.41 to 0.22 with an increasing PTFE volume fraction from 0% to 20%. The addition of PTFE in the PEEK matrix significantly reduced friction of the composite when it slides against a steel counterface. For a PTFE/PEEK composite with less than 15% PTFE (by volume), its friction coefficient decreased monotonically with PTFE volume fraction. 0.5 Ambient Temperature = 23 °C 0.4 µ 0.3 0.2 0.1 0 0 5 10 15 20 Vf (% PTFE) Figure 5.2. Coefficients of friction of neat PEEK and PTFE/PEEK composite as a function of PTFE volume fraction. 59 25 5.2 Experimental Results of Wear The wear losses of neat PEEK, neat PTFE, and PTFE/PEEK composites were determined as a function of sliding time (Figure 5.3). The neat PTFE wear loss was 1.3 mm (in height) in 7 hours of sliding. Wear resistance of neat PEEK was 1.3 mm in height loss in a much longer time (48 hours). All tests for the composite wear lasted 48 hours. The specific wear rates were determined over the second 24-hour period. In the (first 24 hours) running-in period, the wear sample experienced initial thermal expansion due to frictional heating and unstable wear due to incomplete transfer-film development. 1.3 PTFE PEEK Ambient Temperature = 23 °C 1.1 C05 h (mm) 0.9 0.7 0.5 C10 0.3 C15 0.1 C20 -0.1 0 10 20 30 40 t (hour) Figure 5.3. Wear (height) loss as a function of sliding time for PTFE/PEEK composite, neat PEEK, and neat PTFE. 60 The specific wear rate was determined with the procedure described in Section 4.4.3. The (height) wear rate was obtained by taking linear regression of the wear height loss of specimen after the running-in period. Specific wear rates of neat PEEK and PTFE/PEEK composites were shown in Figure 5.4. The specific wear rate of neat PEEK was 1.54×10-5 mm3/Nm. The PTFE/PEEK composites exhibited a monotonic decrease in wear rate with increasing PTFE volume fraction. For the composite with 20% PTFE, the specific wear rate of the composite was 1.20×10-6 mm3/Nm, which was less than 10% of that of neat PEEK. Ambient Temperature = 23 °C αΊ (mm3/Nm) 10-5 10-6 10-7 0 5 10 15 20 25 Vf (% PTFE) Figure 5.4. Specific wear rates of neat PEEK and PTFE/PEEK composites as a function of PTFE volume fraction. 61 5.3 Relationship between Composite Friction and Wear A power-law relationship between friction coefficient (µ) and wear coefficient (K) for metals and non-metals is discussed in [13] with an exponent (n) of 4 for metals and 2 for non-metals, πΎ~π π (5.1) The wear coefficient in [13] is defined as πΎ= π»βΜ , ππ£ (5.2) where H is the material hardness. Comparing with equation 4.8, the difference between wear coefficient and specific wear rate is the material hardness in the numerator. Since determination of hardness of polymeric material depends heavily on test conditions and size of test blocks, the specific wear rate is therefore used in this study to evaluate the wear characteristics of polymeric material. Experimental results show that both the friction coefficient and the specific wear rate PTFE/PEEK composites decreased with an increasing amount of PTFE in the composite. The specific wear rate of the PTFE/PEEK composite was found to relate to the friction coefficient by a power law relationship shown in Figure 5.5. 62 10-4 Ambient Temperature = 23 °C PEEK C05 αΊ (mm3/Nm) 10-5 αΊ = C μβ β=4 C10 C15 C20 10-6 10-7 0.1 1 μ Figure 5.5. Relationship between specific wear rate and friction coefficient of the PTFE/PEEK composite from current experiments. The power-law relationship between specific wear rate and friction coefficient may be expressed as π½ αΊ(ππ ) = πΆ π(ππ ) , where is the exponent, C is a constant and αΊ and (5.3) are the specific wear rate and the friction coefficient, respectively. Note that Eq. 5.3 is similar to the power-law relationship discussed in [13] but the hardness of the material is not involved. In this study, the PTFE volume fraction was in a range from 0 to 20%. In the case of neat PEEK, i.e., , we have 63 π½ αΊππΎ ~ πππΎ . Similarly for the PTFE/PEEK composites, (5.4) , π½ αΊπ ~ ππ . (5.5) Subscripts PK and c denote properties of neat PEEK and the composite, respectively. Taking the ratio of the specific wear rates of neat PEEK and PTFE/PEEK composite, we have ππ π½ αΊc = ( ) ⋅ αΊππΎ . πππΎ (5.6) The equation established a clear relation between friction and wear of neat PEEK and the PTFE/PEEK composite. Consequently, the specific wear rate of the composite can be determined from friction coefficients of the composite and the matrix wear rate. 5.4 Validation with Literature Data For PTFE/PEEK composites, a large amount of experimental data of friction and wear are available in the literature. Note that some [14, 88] only reported the specific wear rate without giving the friction information. The lack of the friction coefficient information provides difficulty in assessing the wear model. However, three sets of experimental results were available [19, 28, 63] to assess the validity of the power-law relationship. While all of the experiments were carried out by sliding PTFE/PEEK composites on smooth steel counterfaces, the sliding speed, normal load, and apparatus used were different. Details of the experimental conditions were summarized in Table 5.1, along with test conditions conducted in this study, for comparison. 64 Table 5.1. Test conditions of PTFE/PEEK friction and wear experiments from the literature. Year Pressure (MPa) Velocity (m/s) Counterface Mat'l Roughness (µm) Test Configuration Lu[28] 1995 1 1 100 Cr 6 0.2-0.3(Ra) Pin-on-Disc Burris[19] 2006 6.25 0.051 304 stainless steel 0.161 (Rq) Reciprocating Vail[63] 2011 6.25 0.051 304 stainless steel 0.15 (Ra) Reciprocating This Study 2018 0.5 1 1018 Carbon Steel 0.1±0.02(Ra) Pin-on-Disc The specific wear rates of PTFE/PEEK composites obtained from [28] as shown in Figure 5.6 and compared with the predictions from Eq. 5.6. Ambient Temperature = 23 °C Eq. (5.6), β=4 αΊ (mm3/Nm) Lu et al. (1995) [28] 0.0 0.1 0.2 0.3 0.4 0.5 Vf (PTFE) Figure 5.6. Experimental results from [28] compared with power-law predictions. 65 In the figure, red circles denote the experimental results and blue squares and lines indicated the solutions determined from Eq. 5.6 with an exponent, . The power-law relationship correctly predicted the decreasing specific wear rate of the composite with increasing amount of PTFE in the composite. Burris and Sawyer examined friction and wear of PTFE/PEEK composite tested on a reciprocal tribometer. The results are also used for validating the power-law relationship. The experimental results and the predictions from Eq. 5.6 were compared in Figure 5.7. Ambient Temperature = 23 °C Eq. (5.6), β=3 αΊ (mm3/Nm) Burris et al (2006) [19] 10-8 0 0.1 0.2 Vf (PTFE) 0.3 0.4 Figure 5.7. Experimental results from [19] compared with power-law predictions. Experimental data obtained by Lu et al. [28] and Burris et al. [19] were compared with the power-law predictions. Good agreement was observed however the power-law 66 model underpredicted the wear rate of the PTFE/PEEK composite with15% PTFE. For the composite with PTFE greater than 30%, specific wear rates of the composites dropped below 10-7 mm3/Nm and the model predictions were also close to the experimental data. Vail et al. [63] performed friction and wear experiments on fibers filled PTFE/PEEK composites with various amounts of PTFE particles. Their test results on wear of PTFE reinforced PEEK were also taken to compare with Eq. 5.6 and shown in Figure 5.8. Ambient Temperature = 23 °C Eq. (5.6), β=3 αΊ (mm3/Nm) Vail et al. (2011) [63] 0 0.05 0.1 0.15 0.2 0.25 Vf (PTFE) Figure 5.8. Experimental results from [63] compared with power-law predictions. The experimental results showed good agreement with the power-law prediction except the case of the composite with 20% PTFE. The wear rate of 20% PTFE/PEEK composite 67 was over 10-6 mm3/Nm while the predictions remained almost the same with that of 15% PTFE/PEEK composite. With all friction and wear results from the relevant literature [19, 28, 63] and from the present research (Figure 5.9), the exponent β of the power law of friction and wear was found to be related to the test methods used. For the pin-on-disk wear test, the β appeared to have a value of 4, whereas the reciprocal friction and wear test gave a β of 3. 10-4 Ambient Temperature = 23 °C αΊ (mm3/Nm) 10-5 β=3 β=4 Test Results 10-6 Vail 2011 [63] Lu 1997 [28] Burris 2006 [19] 10-7 0.1 μ Figure 5.9. Specific wear rates and friction coefficients of PTFE/PEEK composite obtained from experiments with different test methods. 68 1 Chapter 6 Transfer Films 6.1 Nature and Issues of Transfer films in PTFE, PEEK and Their Composite In tribological applications, thermoplastic polymer with low shear strength and surface energy tends to adhere and transfer the material to the high strength and surface energy metal. The layer of polymer on the metal counterface is known as transfer film [89]. With the layer of transfer film developed between the soft polymer and hard metallic counterface, the polymer is prevented from the direct contact with the counterface as illustrated in Figure 2.6. During sliding contact between the PTFE/PEEK polymer composite and the transfer film, the low shear strength of the film may greatly reduce the friction coefficient [90]. The transfer films may also have profound influence on wear of the polymer composite. Transfer films characterized as ‘patchy,’ and ‘nonuniform’ are often related to poor wear resistance, whereas thin and uniform transfer films are associated with the polymer composites having low friction and wear [91–96]. In sliding contact, neat PTFE often forms continuous transfer films with large flake-like debris [49, 51, 56]. Its transfer film is continuously removed and replaced to form large debris during sliding. The layered lattice structure of the PTFE and the low (inter-laminar) shear strength of the PTFE transfer film generally leads to a low friction coefficient, and the rapid removal of PTFE transfer film during sliding results in high wear loss. Neat PEEK tends to form discontinuous transfer films on a metal counterface as discussed in Section 2.3. Though some of small particles of the PEEK may get caught within the valleys of counterface asperities, the total coverage of the PEEK transfer film 69 is not adequate to serve as a protective film to reduce its high friction and wear. Voort et al. [70] observe only small patches of PEEK films on the conterface after sliding. Kalin et al. [97] report that the transfer film coverage of neat PEEK in sliding is less than 27%. Both of them attributed high friction of neat PEEK due to the lack of thin-layer continuous transfer films. The PTFE/PEEK composite, unlike neat PEEK, develops tenacious transfer films on the metal counterface [19, 28, 66]. PEEK polymer filled with PTFE particles has been observed to produce finer wear debris [64]. Accumulation of fine debris on the metal counterface as a thin-layer transfer film results in improvements of friction and wear resistance [19]. Although qualitative description of transfer films of PTFE/PEEK composite suggests that the transfer film deposited on the counterface prevents the direct contact between the composite and the metal surface and reduces both friction and wear, a quantitative evaluation of the transfer film may lead to better understanding of the effect of transfer films on the PTFE/PEEK friction and wear. 6.2 Experiment Methods for Transfer Film Evaluation 6.2.1 Transfer Film Coverage A method for evaluating the transfer film geometry and characterizing the transfer film coverage ratio has been developed in this study. To obtain accurate characterization of its area coverage in sliding contact, the counterface surface was examined under an optical microscope with a non-polarized reflective light source at 25X magnification. A panorama image of the transfer film on the entire counterface surface was obtained by taking multiple micrographs around different locations of the counterface to capture the sectional transfer film images (Figure 6.1, for example). 70 Figure 6.1. Micrographs of transfer film on the counterface (C10 sliding on 1018 carbon steel counterface). The micrographs taken at individual locations were then patched togheter, with the aid of an in-house built computer software, to obtain a high-definition image to determine the total area on the steel counterface that was covered by the transfer film. The highdefinition image was then converted to a gray scale image in which contrast between the transfer film and the bare metal surface was shown by different scales of gray. The intensity of each pixel on the gray-scale image was represented by a digital number (ranging from 0 (black) to 255 (white)). It was possible to establish a threshold intensity (pixel level) to identify the area on the counterface that was covered by the transfer film. 71 Figure 6.2. A method for analyzing images of counterface micrographs to determine the transfer film covered area. A statistical method was used to minimize the error that may be introduced in selecting the threshold intensity level. For example, a small area on the counterface that was totally covered by transfer film was first selected and a histogram of the intensities of the pixels within the area was constructed as shown in Figure 6.2. The mean intensity (μ) and the standard deviation (σ) were then determined. The threshold intensity value that was used to distinguish the area on the counterface that was covered by transfer film from the (uncovered) bare metal surface was then taken as μ+2σ. The fraction of the counterface surface that was covered by the transfer film (i.e., the transfer film areal coverage ratio) was determined with the established threshold intensity. The process was 72 then repeated by choosing a different location on the counterface that was totally covered with the transfer film. Several different locations on the counterface were chosen for analysis and the corresponding fractions of the transfer films coverage were determined. The average value of all the fractions so determined was taken as the average transfer film areal coverage ratio. 6.2.2 Elemental and Compositional Analysis of Transfer Films In addition, elemental and compositional analysis of the transfer film were conducted by X-ray photoelectron spectroscopy (XPS). For the XPS analysis, samples were cleansed by acetone before being placed in a sample chamber. An achromatic Al Kα X-ray source (1486.6 eV) was operated at 350W. The area of measurement, the collection solid angle cone and the take-off angle were set at 800 μm2, 5ºand 45º, respectively. The passing energy of the hemispherical energy analyzer was set at 11.75 eV, which gave a resolution of better than 0.51 eV. The pressure in the vacuum chamber was 5×10-9 torr. Contaminated C 1S (284.9 eV) was employed to calibrate the surface charge effect on binding energy. 6.3 Experimental Observations A recent study of microstructure and function of transfer films formed in sliding friction and wear of PTFE/PEEK composite is reported in [98]. The results of the study indicate that transfer films from PTFE/PEEK composite have a gradient structure with PTFE particles located on the topmost surface of the film and the PEEK polymer lying mainly inside the transfer film. Apparently, the PEEK polymer migrates preferentially first before the developing of the PTFE particles, to the counterface surface. Also, 73 friction reduction is mainly from the PTFE phase lying on the top surface of the thin PEEK film. In the present study, optical micrograph images (Figures 6.3 and 6.4) of wear tracks on a steel counterface obtained from in-house friction tests of neat polymers (PTFE and PEEK) and PTFE/PEEK composite with 20% volume fraction of PTFE. The optical micrograph images shown in Figure 6.3 were obtained with a non-polarized reflective light source. The transfer films formed on the steel counterface from testing the composite (Figure 6.3(b)) and neat PTFE sample (Figure 6.3(c)) can be easily identified in the micrographs. Figure 6.3. Optical (non-polarized) images of steel counterface after wear tests of (a) Neat PEEK polymer, (b) PTFE/PEEK composite (C20) and (c) Neat PTFE polymer. Figure 6.4. Optical (polarized light) images of steel counterface after wear tests of (a) Neat PEEK polymer, (b) PTFE/PEEK composite (C20) and (c) Neat PTFE polymer. 74 The counterface optical images shown in Figure 6.4 were obtained with a polarized reflective light source. PTFE transfer films appeared shiny under the microscope with the polarized light shown in Figure 6.4 (c). Comparing the micrograph image of the counterface (Figure 6.4(b)) from the testing of PTFE/PEEK composite sample with the counterface micrograph of neat PTFE shown in Figure 6.4(c), the same shiny layer appeared on top of the composite transfer films. One may infer that the shiny films on top of the transfer film in Figure 6.4(b) was also PTFE. Similar observations of the PTFE layer were also reported in [98]. Since the PTFE on the top surface of the transfer film behave as a solid-state lubricant, the apparent friction coefficient of transfer films was expected to be similar to that of the neat PTFE polymer. 6.4 Characterization of Transfer Films on Counterface During the sliding contact of PTFE/PEEK composite transfer films were observed to form on the steel counterface surface. The area on the counterface surface covered with the transfer film varied with the volume fraction of PTFE in the composite, from scattered patches at a low PTFE content to almost full coverage at high PTFE content. The transfer film areal coverage ratio (ACR, α) was determined by the method described in Section 6.2 and the results were shown as a function of the PTFE volume fraction (Vf ) in Figure 6.5. The results were also curve-fitted with an error function, πππ πΌ(ππ ) = 2 √π 2 ∫ π −π‘ ππ‘ , (6.1) 0 where is a correlation factor. (The value of 75 equals to 7 is used for the curve fitting.) 1.2 Ambient Temperature = 23 °C 1 α 0.8 0.6 0.4 Eq.6.1, π=7 Measured α 0.2 0 0 0.05 0.1 0.15 Vf 0.2 0.25 0.3 Figure 6.5. Transfer films area coverage ratio for composites with different PTFE volume fractions. Transfer film area coverage ratio (α) was found to increase with the PTFE volume fraction in the PTFE/PEEK composites tested in the study. For the composite with a 20% volume of PTFE, almost the entire counterface was covered by the transfer film. 76 Figure 6.6. XPS scan spectra of virgin and tested steel counterface sliding over the PTFE/PEEK composite. Figure 6.7. XPS C 1s spectra of virgin and tested steel counterface. 77 The results of a compositional analysis of transfer films formed by a PTFE/PEEK composite (analyzed by XPS) are shown in Figures 6.6 and 6.7. Figure 6.6 provided the survey scan of C 1s spectra on a clean counterface before testing and the counterface after 48-hour sliding wear test. On the counterface, after the wear test, a peak of 689 eV was found from the survey scan spectrum. The peak was attributed to the CF2 species. In contrast, no fluorine element peak was found on the virgin counterface. Therefore, one may conclude that during the sliding, PTFE was transferred to the steel counterface. This further confirmed the presence of PTFE in the transfer film together with the previously shown polarized optical micrographs. The survey spectra exhibited decays in intensity of iron and oxygen elements on the tested counterface, suggesting that the thickness of the PTFE film was less than 10 nm, since the detection depth of the XPS was approximately 7 to 10 nm. The C 1s spectra of the virgin surface and tested counterface were shown in Figure 6.7. The C 1s spectrum of the virgin surface exhibited three peaks at 284.8 eV, 285.2eV and 288.2eV. They were attributed to contaminated carbon (C-C), C-O, and C=O, respectively. On the tested counterface, a C-F peak at 292eV was shown. The C-F bond observed on the wear-tested surface must come from the PTFE since no other fluorine source was present in the sliding pair. The XPS elemental study clearly demonstrated the existence of the PTFE in transfer films on the top layer of the transfer film. 78 Chapter 7 Development of Friction and Wear Theories for PTFE/PEEK Composite 7.1 Friction and Mechanical Properties of PTFE/PEEK Composite Mechanical properties (shear and compression yield stresses and elastic modulus) of the neat PEEK and PTFE polymers and PTFE/PEEK composites with different PTFE volume fractions were obtained experimentally in [64, 65, 79]. In Table 7.1 mechanical properties and friction coefficients of the PTFE/PEEK composite are shown. Table 7.1. Mechanical properties and friction coefficients of neat PEEK, PTFE and PTFE/PEEK Composites at room temperature. Vf (PTFE) μ PEEK 0 C05 Yield Stress (MPa) Compression Shear Modulus (GPa) 0.407 143 76 3.87 0.05 0.373 128 59 3.56 C10 0.1 0.300 98 54 3.23 C15 0.15 0.231 85 50 2.97 C20 0.2 0.220 80 49 2.68 PTFE 1 0.208 20 9 0.57 Material In Figure 7.1, friction coefficients of the PTFE/PEEK composites and the neat PEEK are related to their shear, compressive yield stress (cited from [79]) and elastic modulus (E’ obtained from DMA tests). The results indicate that friction coefficient of PTFE/PEEK composite μc (with up to 20% volume fraction of PTFE) may be related to the friction coefficient of neat PEEK (μPK) and its constituent mechanical properties by ππ = ππ π . πππΎ ππΎ 79 (7.1) The PPK and Pc in Eq. (7.1) represent, respectively, mechanical properties (shear yield stresses, compression yield stresses or elastic moduli) of the neat PEEK and the PTFE/PEEK composite. 160 7 Ambient Temperature = 23 °C 140 6 120 5 4 80 3 E (GPa) σy (MPa) 100 60 Neat PEEK 40 2 Compression 1 Shear 20 Modulus 0 0.15 0.2 0.25 0.3 μ 0.35 0.4 0 0.45 Figure 7.1. PTFE/PEEK composite friction coefficients and mechanical properties. Note that subscripts c and PK are used hereafter to denote the quantities associated with PTFE/PEEK composite and neat PEEK polymer. 80 0.45 Ambient Temperature = 23 °C Test Results 0.4 E σyc μ 0.35 τy 0.3 0.25 0.2 0.15 0 0.05 0.1 Vf 0.15 0.2 Figure 7.2. Predicted μc (from Eq. 7.1) and test results on friction coefficients of PTFE/PEEK composite. In Figure 7.2, results of friction experiments on the PTFE/PEEK composite (with up to 20% volume fraction of PTFE) are compared with the predictions from Eq. 7.1 using compression yield stress, shear yield stress and the elastic modulus of the composite (shown in Table 7.1). Good correlations are observed between the test results and the predictions using compressive and shear yield stresses. The predictions with the composite and neat PEEK compressive yield stresses (Eq. 7.1) appear to correlate the best with the test results. The composite friction equation, Eq. 7.1, is established based on the relationships between friction test results and mechanical properties of the PTFE/PEEK composite 81 (with up to 20% volume of PTFE). For the composites with more than 20% volume of PTFE, applicability of Eq. 7.1 for friction coefficient prediction has not been investigated, due to lack of experimental data. A limitation of using Eq. 7.1 to predict friction coefficient of the PTFE/PEEK composite requires, as a priori, shear yield stresses (or compression yield stresses or moduli) of both the composite and the neat PEEK polymer. Also, note that only friction coefficient of the neat PEEK is involved in Eq. 7.1 for predicting the composite friction coefficient. The effect of PTFE on composite friction is implicit through mechanical properties of the composite and friction coefficient of PTFE does not appear explicitly in Eq. 7.1. 7.2 Solid Film Lubrication and Associated Models When a soft lubricating film is introduced as an interphase between a steel counterface and a sliding PEEK polymer, the apparent friction coefficient between the neat PEEK and the counterface is governed by the plastic flow of the lubricating film as ππ = ππ ⋅π , πππΎ ππΎ (7.2) where μl and τl are friction coefficient and shear yield stress of the lubricating film, respectively, and τPK is shear yield stress of neat PEEK polymer. Two special cases are considered below. If the lubricating film has the same friction and mechanical properties as those of the PTFE/PEEK composite, then μl or the composite friction coefficient (μc) may be expressed in terms of the neat PEEK polymer friction coefficient as π π = ππ = ππ ⋅π , πππΎ ππΎ 82 (7.3) where τc is shear yield stress of the composite. If the sliding PEEK polymer is replaced with the PTFE/PEEK composite, then τPK and μPK in Eq. 7.3 would take on the values of τc and μc and it simply becomes an identity ππ = ππ ⋅π . ππ π (7.4) Hence, friction coefficient behavior of a PTFE/PEEK composite may not be directly determined with the conventional solid film lubrication theory [13]. Further, according to the experimental relationship obtained in Section 7.1, the composite shear yield stress (τc) and neat PEEK shear yield stress (τPK) are used as terms Pc and PPK in Eq. 7.1. Substituting τc and τPK in Eq.7.1, it becomes identical to Eq. 7.3 for the PTFE/PEEK composites with up to 20% volume of PTFE. This implies that if shear yielding is the controlling mechanism of friction, then friction behavior of PTFE/PEEK composite (with up to 20% volume of PTFE) would be similar to that of neat PEEK polymer sliding on thin films of PTFE/PEEK composite located on the top surface of the steel counterface. The friction coefficient of the composite, μc, was in fact originated from the shear of the transfer film. Similar to the friction coefficient of lubricating film, μl, from the solid-film lubrication theory, transfer-film friction coefficient, μTF, was defined associated with the shear of the composite transfer film and approximated as π ππΉ = ππ ⋅π . πππΎ ππΎ 83 (7.5) 7.3 Friction Involving Transfer Films To account for the effect of transfer films on friction of PTFE/PEEK composite in sliding contact, the contact area (test pin cross-section area) between the composite sample and the steel counterface is assumed to consist of two distinct areas – one with and the other without the transfer film coverage. Friction of the composite in the contact area with transfer films is different from that of the area lacking the transfer film coverage. Further, the contact area without the transfer film coverage is divided into two parts: the area of PTFE covering the counterface and the other of neat PEEK on the steel counterface. Modifications are made to the linear rule mixtures (LROM) and inverse rule of mixtures (IROM) to incorporate the effect of transfer films on the overall friction of the PTFE/PEEK composite. Details of the modifications are given in Appendix B. Two expressions are obtained for predicting the friction coefficient (μc) of the PTFE/PEEK composite. With the modified LROM, one has ππ = πΌπ ππΉ + (1 − πΌ)ππ πππ + (1 − πΌ)(1 − ππ )πππΎ , (7.6) where the first term on the right-hand side of Eq. 7.6 is the contribution of transfer films to the friction coefficient of the composite. Similarly with the modified IROM, one obtains ππ = 1 (1 − πΌ)ππ (1 − πΌ)(1 − ππ ) πΌ + + π ππΉ πππ πππΎ 84 . (7.7) In Eqs. 7.6 and 7.7, α, Vf, µTF and µPK are the transfer film areal coverage ratio (ACR), the volume fraction of PTFE, the transfer film friction coefficient and friction coefficient of neat PTFE polymer, respectively. Note that the areal coverage ratio (ACR) is a function of Vf as shown in Fig. 6.5. In the following, two different values are assumed for the apparent friction coefficient (µTF) of transfer films. First, the transfer films are assumed to behave as a soft lubricant with a friction coefficient given by Eq. 7.5, i.e., π ππΉ = ππ ⋅π . πππΎ ππΎ (7.5) The PTFE/PEEK composite coefficient of friction, µc, from Eqs. 7.6 and 7.7, are then obtained as ππ = πΌ ππ π + (1 − πΌ)ππ πππ + (1 − πΌ)(1 − ππ )πππΎ πππΎ ππΎ (7.8) or ππ = 1 (1 − πΌ)ππ (1 − πΌ)(1 − ππ ) πΌ + + ππ πππ πππΎ π ππΎ π , (7.9) ππΎ where μPT is the friction coefficient of neat PTFE. Second, the PTFE material on top of the transfer films, as shown in Figure 6.4, are assumed to behave as a solid-state lubricant and the friction coefficient of transfer films is the same as that of the neat PTFE, i.e., 85 π ππΉ = πππ . (7.10) Incorporating Eq. 7.10 into Eqs. 7.6 and 7.7, the friction coefficient of the PTFE/PEEK composite may be expressed as ππ = [πΌ + (1 − πΌ)ππ ]πππ + (1 − πΌ)(1 − ππ )πππΎ (7.11) or ππ = 1 . πΌ + (1 − πΌ)ππ (1 − πΌ)(1 − ππ ) + πππ πππΎ (7.12) The PTFE/PEEK composite friction coefficient determined by Eqs. 7.8 and 7.9 are compared with the experimental results obtained in the study in Figure 7.3. The friction coefficient predictions by Eqns. 7.11 and 7.12 are also shown in the figure. For PTFE/PEEK composite with a low PTFE content (≤ 10% by volume), the composite friction solutions by Eqs. 7.8 and 7.9 are closer to the test results than those predicted by Eqs. 7.11 and 7.12. However, for the PTFE/PEEK composite with a high PTFE volume fraction (more than 10%), the solution obtained by Eqs. 7.11 and 7.12 are closer to the test results. Overall, the theoretical solution based on the linear rule-of-mixtures with the transfer film having a friction coefficient of neat PTFE yields better results closer to the experimental data. 86 0.45 Ambient Temperature = 23 °C 0.4 μ 0.35 0.3 0.25 Eq. 7.8 Eq. 7.9 Eq. 7.11 Eq. 7.12 Experimental results 0.2 0.15 0 0.05 0.1 Vf (PTFE) 0.15 0.2 Figure 7.3. Comparison of PTFE/PEEK composite friction coefficient predictions with experimental results. Equations 7.8 and 7.9 are obtained, based on modification of the LROM and IROM for the PTFE/PEEK composite. The transfer films are considered to have friction behavior similar to a solid (composite) lubricating film. Since the transfer-film ACR (α) is required, in addition to the input of composite and PEEK matrix shear yield strengths, Eqs. 7.8 and 7.9 are more involved than Eq. 7.1. As the PTFE content in the composite increased beyond 20% (by volume), the transfer-film ACR (α) rapidly approched unity (Figure 6.5), and Eqs. 7.8 and 7.9 revert to the one of the three cases represented by Eq. 7.1 (i.e., the case based on the composite and the neat PEEK shear yield stresses). 87 The friction solutions, Eqs. 7.11 and 7.12 are also derived based on modifications of the LROM and IROM and may be convenient (compared with Eqs. 7.8 and 7.9) to use since no composite mechanical properties are required. However, the transfer-film ACR (α) is still needed for the composite friction coefficient determination. From microscopic observations, surface compositional analysis and the literature [98], the transfer-film friction coefficient is chosen to be the friction coefficient of neat PTFE. For the composite lubricated with more than 10% PTFE, the PTFE-film lubrication-based model yields the best theoretical results when compared with experimental data. For the PTFE volume fraction below 10%, the model underestimated the friction coefficient of the composite due to the insufficient amount of PTFE retained on top of the transfer film to form a PTFE lubricating layer. Thus, the solid film lubrication theories with a composite transfer film lead to analytical results closer to the experimental data. 7.4 Wear with Transfer Films Wear of the PTFE/PEEK composite is reduced due to the synergetic effect of the transfer film lubrication [26]. The large plastic flow and low surface energy of the transfer film appear to reduce the friction during dry sliding. Wear, a form of progressive material failure, is related to cohesive energy of the polymer material. With less frictional force, the probability of material damage is lowered, leading to less wear. For a PTFE/PEEK composite (with up to 20% of PTFE by volume), this behavior was shown by a power-law relationship (between the friction and wear) as discussed in Section 5.3. Since friction of PTFE/PEEK composite is successfully modeled by the modified LROM and IROM with considerations of transfer films, wear of the PTFE/PEEK composite 88 material may also be modeled with a combination of transfer film lubrication friction theory and the power law relationship. Considering the LROM and IROM with composite transfer films (Eqs. 7.8 and 7.9), specific wear rate, αΊ, of PTFE/PEEK composite (with up to 20% PTFE by volume) may be determined as αΊπ = {(1 − πΌ) [ππ ( πππ ππ π½ − 1) + 1] + πΌ } αΊππΎ μππΎ πππΎ (7.13) and π½ 1 αΊπ = { π } αΊππΎ . π πΌ πππΎ + (1 − πΌ) [ππ ( πππΎ − 1) + 1] π (7.14) ππ Similarly, if the specific wear rate of PTFE/PEEK composite is modeled with the PTFE lubrication theory (Eqs. 7.11 and 7.12), are obtained αΊπ = {[πΌ + (1 − πΌ)ππ ] π½ πππ + (1 − πΌ)(1 − ππ )} αΊππΎ πππΎ (7.15) and π½ 1 αΊπ = { } αΊππΎ . πππΎ [πΌ + (1 − πΌ)ππ ] + (1 − πΌ)(1 − ππ ) πππ (7.16) The PTFE/PEEK composite specific wear rate obtained by Eqs. 7.13 and 7.14 are shown with the experimental data in Figure 7.4. The specific wear rate given by Eqns. 7.15 and 7.16 are also given in the figure. For the composite with a 5% PTFE volume content, the solutions for Eqs. 7.13 and 7.14 are found closer to the experimental results than the 89 solutions determined by Eqs. 7.15 and 7.16. However, for the composite with more than 10% PTFE (by volume), the predictions obtained by Eqs. 7.15 and 7.16 exhibit better agreements with experimental data. 10-4 Eq. 7.13 Eq. 7.14 Eq. 7.15 Eq. 7.16 Experimental results Ambient Temperature = 23 °C αΊ (mm3/Nm) 10-5 10-6 10-7 0 0.05 0.1 Vf 0.15 0.2 Figure 7.4. Comparison of specific wear rate solutions with experimental results (β=4). The wear of PTFE/PEEK composite depends highly on its friction behavior. With the presence of transfer films, adhesion between the composite test sample and the counterface was reduced, especially with a layer of PTFE deposited on the top of the transfer film. The reduction in adhesion resulted in low friction and wear at the same time. This synergistic effect for PTFE lubricated PEEK composite is successfully modeled by the modified LROM and IROM including the effect of lubricating transfer films. 90 Chapter 8 Elevated Temperature Friction and Wear of PTFE/PEEK Composite 8.1 Elevated Temperature Friction and Wear Experimental Results Tribological tests of neat PEEK, neat PTFE and PTFE/PEEK composites (C10, C15, and C20) were carried out at elevated temperature up to 200 °C. Test duration for the neat PEEK and composite samples was 48 hours while for neat PTFE it was 8 hours due to high wear rate. The test sliding speed and applied pressure in all tests were 1m/s and 0.5 MPa, respectively. The experimental matrix of elevated temperature friction and wear tests is shown in Table 4.3. During each test, contact surface temperature was controlled to be a constant. Friction force and wear loss were measured and recorded every 2 minutes. 8.1.1 Friction The results of sliding-contact friction experiments of neat PEEK, neat PTFE, C10 and C15 composite at 60 °C and 200 °C were shown in Figure 8.1. At 200°C, the neat PEEK reached the highest friction coefficient, 0.6. The friction coefficients of C10, C15, C20 and neat PTFE were found to be 0.3, 0.24, 0.21 and 0.14, respectively. At 60 °C, the friction coefficient of neat PEEK dropped from 0.6 to 0.4 while the C15 and neat PTFE exhibited friction coefficients of around 0.27. The friction coefficient of C10 was about 0.3 at 60 °C, higher than these of neat PTFE and C15. The elevated temperature friction and wear experimental results were summarized in Table 8.1. 91 0.7 PEEK @ 200 °C 0.6 0.5 μ PEEK @ 60 °C 0.4 C10 @ 200 °C 0.3 PTFE @ 60 °C C10 @ 60 °C C15 @ 60 °C 0.2 C15 @ 200 °C PTFE @ 200 °C 0.1 0 0 10 20 t (hour) 30 40 50 Figure 8.1. Friction coefficients during sliding wear of neat PEEK, neat PTFE, C10 and C15 composites at 60 and 200 °C. Table 8.1 Results of friction and wear experiments of neat PEEK, neat PTFE and the PTFE/PEEK composites in sliding contact at different temperatures. PEEK C10 C15 C20 PTFE T (°C) μ αΊ (mm3/Nm) μ αΊ (mm3/Nm) μ αΊ (mm3/Nm) μ αΊ (mm3/Nm) μ αΊ (mm3/Nm) 60 0.42 1.03×10-5 0.30 3.08×10-6 0.27 1.95×10-6 0.23 1.06×10-6 0.20 6.65×10-5 100 0.42 1.38×10-5 0.31 3.18×10-6 0.27 2.45×10-6 0.24 1.21×10-6 0.20 2.04×10-5 125 0.42 1.15×10-5 0.26 1.36×10-6 0.22 8.65×10-7 - - 0.14 1.80×10-5 150 0.43 1.24×10-5 0.26 1.68×10-6 0.22 7.36×10-7 0.20 5.09×10-7 0.14 1.88×10-5 175 0.63 1.55×10-6 0.29 6.05×10-7 0.23 4.68×10-7 - - 0.15 2.01×10-5 200 0.63 1.40×10-6 0.30 5.16×10-7 0.24 4.75×10-7 0.21 3.79×10-7 0.14 2.23×10-5 92 As shown in the Table, friction coefficient of neat PEEK increased significantly above at the PEEK glass transition temperature while fiction coefficients of C10 and C15 dropped between the PTFE α phase transition (116 °C) and PEEK glass transition (152 °C). A monotonic decreasing trend of neat PTFE friction with increasing temperature was observed. Similar results were reported in [99]. The effect of temperature on PTFE friction may be related to its viscoelastic behavior, which depended on temperature and strain rate. The temperature-dependent friction behavior of PTFE may be described with a modified Arrhenius equation in [51]. 8.1.2 Wear Height losses during sliding contact of neat PEEK, neat PTFE, C10 and C15 composites at different temperatures were obtained and shown in Figure 8.2. the negative height loss was observed at the beginning of the sliding test due to controlled (prescribed) and frictional heating. The sliding contact between the steel counterface and polymer pins resulted in smooth contact surfaces after the initial wear of the polymer pin and followed by a steady-state wear. Similar to the room temperature experiment, the test period prior to the steady-state wear was the “running-in” phase. From Figure 8.2, neat PEEK and PTFE/PEEK composites, wear rates reached steady state after 24 hours into the tests. Hence wear in the first 24-hour running-in time was not considered in the subsequent analysis of the specific wear rate of the test material. 93 2.5 PTFE @ 60 °C 2 PTFE @ 200 °C h (mm) 1.5 PEEK @ 60 °C 1 C10 @ 60 °C 0.5 C10 @ 200 °C C15 @ 60 °C 0 C15 @ 200 °C 24h -0.5 0 10 20 t (hour) 30 PEEK @ 200 °C 40 50 Figure 8.2. Wear of PEEK, PTFE, C10 and C15 composite at 60°C and 200°C. In Table 8.1, the specific wear rate of neat PEEK was found around 10-5 mm3/Nm below its glass transition temperature, Tg (152 °C). At elevated temperature above Tg, the wear rate of neat PEEK was reduced by one order of magnitude to 1.55×10-6 mm3/Nm. The specific wear rate of the neat PTFE decreased from 6.65×10-5 mm3/Nm to 2×10-5 mm3/Nm during the temperature increase from 60°C to 100 °C. Above 100 °C, the specific wear rate remained almost constant. Wear rates of all composites decreased with increasing temperature. For example, the C15 composite had a specific wear rate of 2×10-6 mm3/Nm at 60 °C and 5×10-7 mm3/Nm at 200 °C. 94 8.2 Characteristics of Friction and Wear at Elevated Temperature At the end of a sliding test of the neat PEEK at 60 °C, only a limited amount of PEEK residual was observed on the steel counterface. As shown in Figure 8.3 (c), only a small amount of PEEK accumulated along the polishing marks on the counterface and no patchy transfer films were observed. The sample worn surface was smooth with slight scratching bands observed along the sliding direction. The scratching marks may result from abrasion by contaminated hard particles or PEEK wear debris. Small (scale-like) debris were produced during the sliding test as shown in Figures 8.3 (b) and (d). No polymer drawing nor significant plastic flow was observed at the leading and trailing edges of the PEEK sample pin after each test. At elevated temperature (200°C), neat PEEK exhibited different behavior during the sliding experiment. As shown in Figures 8.4 (a) and (c), neat PEEK formed continuous transfer films on the steel counterface. Above its glass transition temperature, PEEK became viscoplastic and had much lower yield stress with large elongation at fracture (as discussed in Section 4.2). Observed transfer films and the worn pin surface indicated PEEK had large plastic flow at 200 °C during the sliding experiment. In addition, the polymer extrusion and drawing at the trailing edge of the pin further confirmed the plastic flow of neat PEEK when sliding at elevated temperature. The PEEK friction had a combined effect of adhesion at the contact interface and drawing of polymer chains. 95 Figure 8.3. (a) Steel counterface (b) sample worn surface and SEM images of (c) steel counterface and (d) worn surface of neat PEEK after friction and wear sliding test at 60°C. Figure 8.4. (a) Steel counterface, (b) sample worn surface and SEM images of (c) steel counterface and (d) worn surface of neat PEEK after firciton and wear sliding test at 200°C. 96 The worn sample surface and steel counterface characteristics of the PTFE/PEEK composites (C15, for example) after sliding at 60°C are shown in Figure 8.5. Thin transfer films were observed on the steel counterface after sliding. On the sample worn surface, PTFE particles (in light color) were seen embedded in the PEEK matrix. The worn surface was mostly smooth with minimal scratches. Wear debris of the composite were observed in fine power form. No polymer drawing was seen at the trailing edge of the composite sample at the low temperature (60°C). Figure 8.5. (a) steel Steel counterface, (b) sample worn surface and SEM images of (c) steel counterface and (d) worn surface of C15 composite after friction and wear sliding tests at 60°C. 97 Elevated temperature tribological characteristics of the PTFE/PEEK composites were shown in Figure 8.6 (using C15 as an example). Transfer films were found on the steel counterface similar to the one observed on the neat PEEK sample during a sliding experiment. This indicated large plastic flow and transfer of the PEEK matrix. Polymer extrusion and drawing were also observed at the trailing edge of the composite sample pin, though to a less extent when compared with that observed for neat PEEK. This may result from significant reduction in sliding friction due to PTFE lubrication. Figure 8.6. (a) Steel counterface, (b) sample worn surface and SEM images of (c) steel counterface and (d) worn surface of C15 composite after friction and wear sliding tests at 200°C. 98 At 60°C, neat PTFE showed significant wear which may be attributed to rapid destruction of banded crystalline PTFE structure and the large amount of material transfer to the counterface. As shown in Figure 8.7 (c), a large amount of PTFE transfer films (in dark color) was observed on the steel counterface. At the trailing edge of PTFE pins, thin strips of PTFE were drawn from the bulk (Figure 8.7 (b)). The large amount of PTFE debris accumulated at the leading edge of PTFE pins was due to breakage of PTFE transfer films. At 200 °C, the similar drawing effect was seen on PTFE sliding pins. However, the optical micrograph of the steel counterface showed thin and uniform transfer films, which may be responsible for decreases of wear at elevated temperature. Figure 8.7. Steel counterface, (a) and (d); PTFE sample worn surface, (b) and (e), and micrographs of counterface, (c) and (f). (a), (b) and (c) from tests at 60°C, and (d), (e) and (f), from 200°C. 99 8.3 Relationship between Friction and Wear at Elevated Temperature The power-law relationship between friction and specific wear rate of neat PEEK and PTFE/PEEK composites obtained in Chapter 5 is found also applicable to the tribological results at elevated temperature, i.e., ππ (π) π½(π) αΊc (T) = ( ) ⋅ αΊππΎ (π) , πππΎ (π) (8.1) where αΊc(T), αΊPK(T), μc(T) and μPK(T) are specific wear rates and friction coefficients of the composite and neat PEEK, respectively. Note that the exponent, β, is a function of contact-surface temperature depending on the temperature below or above the glass transition temperature of the matrix material. Figure 8.8, the power law relationship between friction and wear of neat PEEK and the composites are shown for the cases of 100 °C, 125 °C, 150 °C, 175 °C and 200 °C. Below the glass transition temperature (152°C) of PEEK, the value of β is close to 4 and above the Tg it is about 1. Experimental results of friction and wear of neat PEEK and C10, C15 and C20 and the values of β at different temperatures are given in Table 8.2. 100 101 0.43 0.63 0.63 150 175 200 0.26 0.29 0.30 1.24×10-5 1.55×10-6 1.40×10-6 0.31 1.38×10-5 0.26 0.30 1.03×10-5 1.15×10-5 μ PEEK αΊ (mm3/Nm) 5.16×10-7 6.05×10-7 1.68×10-6 1.36×10-6 3.18×10-6 3.08×10-6 C10 αΊ exp. (mm3/Nm) 6.07×10-7 6.13×10-7 1.24×10-7 1.81×10-7 4.00×10-6 2.93×10-6 αΊc theo. (mm3/Nm) * Value of β is determined for each individual temperature. 0.42 0.42 100 125 0.42 μ 60 T (°C) 0.24 0.23 0.22 0.22 0.27 0.27 μ 4.75×10-7 4.68×10-7 7.36×10-7 8.65×10-7 2.45×10-6 1.95×10-6 C15 αΊ exp. (mm3/Nm) 4.79×10-7 4.70×10-7 6.65×10-7 8.42×10-7 2.49×10-6 1.83×10-6 αΊc theo. (mm3/Nm) 0.21 0.20 0.24 0.23 μ 3.79×10-7 5.09×10-7 1.21×10-6 1.06×10-6 C20 αΊ exp. (mm3/Nm) 4.21×10-7 4.18×10-7 1.46×10-6 9.86×10-7 αΊc theo. (mm3/Nm) 1.1 1.2 3.8 3.9 3.8 3.9 β* Table 8.2. Experimental and theoretical predictions of friction and wear of neat PEEK and PTFE/PEEK composites, and the exponent β in power law relation. 10-4 αΊ (mm3/Nm) 10-5 αΊc=C1µβ β≈4 100 °C 125 °C 10-6 150 °C 10-7 0.1 μ (a) 1 10-5 175 °C αΊ (mm3/Nm) 200 °C 10-6 αΊc=C2µβ β≈1 10-7 0.1 μ (b) 1 Figure 8.8. Power law relationship between friction and wear at various temperatures, (a) below Tg of PEEK matrix (152 °C), (b) above Tg of PEEK matrix. 102 The specific wear rates, αΊc, obtained by the power-law relationship (Eq. 8.1) exhibited good agreement with experimental data at all temperatures. The values of β at different contact surface temperatures were determined and shown in Figure 8.9. Also shown in Figure 8.9 is the storage modulus of neat PEEK. Variation of the exponent β with temperature is seen to follow that of the neat PEEK storage modulus. So far, the value of β is based on experimental observations. The physical meaning of β is not determined at this time but is likely related to the mechanical properties of the PTFE/PEEK composites as shown in Figure 8.9. For temperature below the PEEK glass transition temperature, β is probably associated with the elastic properties of the PTFE/PEEK composites and for temperature above the PEEK glass transition temperature, β in the power-law relationship may be related to viscoplastic behavior of the composite. Exact relationship between β and composite elastic and viscoplastic behavior is still unclear at this time and worth further study. Friction and wear of the PEEK polymer at T>Tg behaved differently from that of below Tg due to viscoelastic/plastic transition of the PEEK material. 8.4 Elevated Temperature Friction Theory Experimental observations of the steel counterface after sliding wear test of the PTFE/PEEK composite showed that the counterface was fully covered with transfer films at temperature above Tg, independent of the volume fraction of PTFE in the composite. Hence the friction theory (Chapter 7) based on transfer film area coverage ratio for room temperature is not adequate to address the friction of composites with different amounts of PTFE above Tg. In this chapter, a new friction theory is proposed to address the PTFE/PEEK composite tribology at all temperatures. 103 5 4.5×109 4.0×109 4 3.5×109 β 2.5×109 2.0×109 2 E' (Pa) 3.0×109 3 1.5×109 β 1 1.0×109 E' of PEEK 5.0×108 0 0 50 100 150 200 0 250 T ( °C ) Figure 8.9. Values of β in the power-law relationship and the storage modulus of neat PEEK. An analytical theory that may be adequately capable of studying and predicting friction of PTFE/PEEK composite with a wide range of PTFE volume fraction would be difficult due to the complex mechanisms involved in friction and wear of the composites. The composite microstructure and the PTFE lattice structure [19, 47, 100, 101] have been noted to lead to friction and wear mechanisms that vary with the amount of PTFE in the composite. In addition, the method and conditions used in a friction test could also influence the friction coefficient of the composite. In view of the many factors that could 104 affect friction behavior of PTFE/PEEK composite, to develop a suitable friction theory for the PTFE/PEEK composites with a PTFE volume content ranging from 0% to 100%, proper assumptions need to be made on the composite microstructures and apparent friction of PTFE phase in the composite as well as the friction test method. Consider a PTFE/PEEK composite with low PTFE volume fraction in sliding friction and wear. The PTFE particles that broke loose or detached from the PEEK matrix during sliding functioned as solid-state lubricants. The lubricating PTFE particles were ∗ assumed to have a friction coefficient (πππ ) lower than that of the neat PTFE (π Μ Μ Μ Μ Μ ) ππ due to (a) unique lattice structure of the PTFE crystallite as shown in Figure 8.10 [47], (b) interlayer sliding of crystalline slices as shown in Figures 8.10 and 8.11 [47, 98], and (c) the lack of sufficient bulk PTFE materials constraining the exfoliation of the crystalline slices (Figure 8.1) [47]. Figure 8.10. PTFE crystallite structure (following the illustrations in [47]). 105 Figure 8.11. Schematic PTFE sliding mechanisms (following the illustrations in [98] and [47]). The apparent friction coefficient of the PTFE particles in the composite is approximated by ∗ (π) πππ = π Μ Μ Μ Μ Μ (T) πππ where composite, and for (ππ ≤ ππΆ ), (8.2) is the apparent friction coefficient of the lubricating PTFE particles in the is the friction coefficient of the neat PTFE. The in Eq. 8.2 is a parameter whose magnitude depends on the test system used and the PTFE content in the composite. The physical meaning of ξ is a measure of the degrees of difficulty for the PTFE particle within the PEEK matrix to exfoliate and is a reflection of the combined effect of PTFE morphology, composite microstructure and testing method. The size of PTFE crystals (micron level) may not significantly affect the value of ξ in view of the 106 size of crystalline slices (nanometer level) involved in the sliding mechanism and microstructure of the PTFE/PEEK composite with low volume fraction of PTFE. However, further work is needed to provide a clear understanding of the parameter ξ . Based on the discussions above, the ξ is assumed unchanged below a critical PTFE volume fraction (φc) in the composite. As the volume fraction of PTFE in the composite increased above the critical volume fraction (φc), the apparent friction coefficient of the PTFE in the composite became high due to the constraint on exfoliation of crystalline slices of PTFE exerted by the increasing amount of surrounding PTFE material. In this study, the apparent friction coefficient of PTFE is assumed to increase linearly with the PTFE volume fraction until it becomes the same as the friction coefficient of the neat PTFE (i.e., 100% PTFE volume fraction in composite), ∗ (π) πππ = π Μ Μ Μ Μ Μ (T) πππ + Μ Μ Μ Μ Μ π ππ (π) (1 − π) (ππ − ππ ) (1 − ππ ) for ππ ≥ ππΆ . (8.3) Hence in this engineering friction model, bilinear friction is approximated by the lubricating PTFE during sliding. The values of φc and ξ chosen to reflect the influences of the composite microstructure, morphology of PTFE lubricant, and experimental conditions as shown in Figure 8.12. 107 μ*PT μPT ξ μPT 0 φc 0.2 Figure 8.12. Approximate relation of 0.4 0.6 Vf (PTFE) 0.8 1 and Vf of PTFE in PTFE/PEEK composite. Based on the PTFE/PEEK composite microstructure considerations during slidng contact of the composites with different PTFE volume fractions, the value of φc is chosen to be 0.15. As discussed previously, the parameter ξ is assumed to vary with the type of tribo-system in sliding experiment. For the test with a pin-on-disk or block-on-ring (rotating), type tribo-tester, ξ is assumed to be 0.4, whereas for the tests with pin-on-plate (reciprocating) type tribo-tester, the ξ is smaller and assumed as to 0.2. The smaller ξ for the pin-on-plate type reciprocating test accounts for the larger amount of PTFE debris being retained and accumulated on wear tracks during a sliding test than that are retained on wear tracks during a spinning pin-on-disk (or the block-on-ring) test. Based on the uniform-shear friction model given in Chapter 5, the friction coefficient of the PTFE/PEEK composite (μc) may be expressed by ππ 1 − ππ 1 = ∗ + , ππ (π) πππ (π) πππΎ (π) 108 (8.4) where Vf is the volume fraction of PTFE phase in the composite. Eq. 8.4 is applicable for all PTFE volume fractions in PTFE/PEEK composites. Note that only the friction coefficients Μ Μ Μ Μ Μ πππ and πππΎ of neat PTFE, neat PEEK and the volume fraction Vf of PTFE are needed for determining the composite friction coefficient ππ . The validity of the theory for PTFE/PEEK composite friction will be checked against the experimental results on PTFE/PEEK composites obtained at room and elevated temperatures as well as the test data reported in the literature. 8.4.1 Validation of Friction Theory with Room Temperature Experiments In Figures 8.13 to 8.16, friction coefficients of PTFE/PEEK composites are determined and compared with experimental results [19, 63, 98]. The relevant parameters used in Eq. 8.4 for PTFE/PEEK friction coefficient predictions are summarized in Table 8.3. Good agreement between the predictions and the experimental results is observed for the full range of PTFE volume content (0% to 100% by volume) at room temperature. The predictions by Eq. 8.4 capture the characteristic changes in the PTFE/PEEK composite friction coefficient with increasing PTFE volume content. Table 8.3. Parameters used for Eq. 8.4 to determine composite friction coefficient. Reference Test Type* This study Pin-on-Disk 0.208 0.407 0.4 0.15 Vail et al. [63] Pin-on-Plate 0.135 0.37 0.2 0.15 Burris et al. [19] Pin-on-Plate 0.135 0.363 0.2 0.15 Onodera et al. [98] Pin-on-Disk 0.208 0.288 0.4 0.15 * With different counterface materials. 109 Figure 8.13. Comparison between friction theory (Eq. 8.4) and experimental results obtained in this study. Figure 8.14. Comparison between friction theory (Eq. 8.4) and experimental results from [63]. 110 Figure 8.15. Comparison between friction theory (Eq. 8.4) and experimental results from [19]. Figure 8.16. Comparison between friction theory (Eq. 8.4) and experimental results from [98]. 111 8.4.2 Validation of Friction Theory with Elevated Temperature Experiments Friction and wear experiments on PTFE/PEEK composite conducted from 60 °C to 200 °C were compared with the theoretical predictions (Eq. 8.4). The measured friction coefficients at different temperatures are found in good agreement with theoretical predictions (Figure 8.17). At elevated temperature, neat PTFE and PTFE/PEEK composites exhibited a reduction in friction around 116 °C. This may attribute to the α phase transition of the PTFE. DMA results discussed in Section 4.22 (b) showed that above the α transition, loss modulus of the PTFE decreased. This suggested a lower energy dissipation and a higher molecular mobility of PTFE polymer chains. Similar observations of the temperature dependent friction behavior of PTFE are also reported in [102]. Further increase in temperature above the glass transition temperature of neat PEEK, resulted in higher friction coefficients of PTFE/PEEK composite. Friction of PTFE/PEEK composite at elevated temperature was affected by both thermal transitions of the PTFE and PEEK phase. 112 0.7 Experimental Eq. 8.4, ξ=0.4 PEEK Test Results PTFE Test Results 0.6 PEEK 0.5 μ 0.4 0.3 C10 C15 0.2 C20 PTFE 0.1 0 25 50 75 100 125 T (°C) 150 175 200 225 Figure 8.17. Theoretical and experimental friction coefficients of neat PEEK, neat PTFE and PTFE/PEEK composites at elevated temperature. 8.5 Elevated Temperature Wear Theory A new wear theory for PTFE/PEEK composite has also been developed in this research based on the power-law relationship of friction coefficient. Combining Eqs. 8.1 and 8.4, we have αΊc (T) = αΊPK (π) π½(π) ∗ πππ (π) − πππΎ (π) [ππ ( ) + 1] πππΎ (π) 113 . (8.5) This equation provides a straightforward approach to determine the specific wear rate of the PTFE/PEEK composite at elevated temperature provided that the friction and wear properties of neat PEEK and PTFE are known. Input data to wear equation (Eq. 8.5) and specific wear rate predictions for C10, C15 and C20 composite as shown in Table 8.4. For simplicity, two values of the exponent β, 3.8 and 1.2, are used in the predictions, one for below and one for above the Tg of PEEK. Different friction and wear characteristics of the PTFE/PEEK composite were observed below and above the glass transition of PEEK. The viscoelastic/plastic transition of the PEEK matrix altered the friction and wear mechanisms of the composite. The change in friction mechanisms resulted in different values of β. The theoretical predictions and experimental data obtained from the current experiments (discussed in Section 8.1) are given in Figure 8.18. T (°C) Table 8.4. Parameters used for Eq. 8.5 to determine composite wear rate. Input Parameter Predictions (Eq. 8.5) αΊ PEEK μ PEEK μ PTFE β αΊ C10 αΊ C15 αΊ C20 60 1.03×10-5 0.42 0.20 3.8 2.69×10-6 1.58×10-6 1.12×10-6 100 1.38×10-5 0.42 0.20 3.8 3.37×10-6 1.95×10-6 1.37×10-6 125 1.15×10-5 0.42 0.14 3.8 1.67×10-6 8.38×10-7 5.54×10-7 150 1.24×10-5 0.43 0.14 3.8 1.45×10-6 6.92×10-7 4.48×10-7 175 1.55×10-6 0.63 0.15 1.2 6.93×10-7 5.33×10-7 4.29×10-7 200 1.40×10-6 0.63 0.14 1.2 5.97×10-7 4.55×10-7 3.64×10-7 * Units of specific wear rate, αΊ, are mm3/Nm. 114 10-5 Experimental αΊ (mm3/Nm) Eq. 8.5, ξ=0.4 10-6 C10 C15 C20 10-7 50 75 100 125 150 T (°C) 175 200 225 Figure 8.18. Specific wear rates of PTFE/PEEK composites (C10, C15, and C20) and theoretical predictions. With known friction and wear properties of the neat PEEK and the neat PTFE, the wear theory developed in this study successfully predicts specific wear rates of the PTFE/PEEK composites at all temperatures. The relationship established between friction and wear of PTFE/PEEK composite reveals the dependence of wear on friction-induced interface shear during sliding. For the composites, transfer films introduced by the PTFE solid-state lubricant reduce the friction resistance. The reduction in friction then lowers the shear stress along the contact surface, which in turn decrease the material damage and loss. Friction and wear mechanisms of PTFE/PEEK at low and elevated temperatures are 115 complicated and different. The friction and wear theories proposed in this study are based on experimental study and observations. Friction and wear mechanisms of PTFE/PEEK composite at elevated temperature are discussed in detail in the next chapter. 116 Chapter 9 Mechanisms of Friction and Wear of PTFE/PEEK Composite at Elevated Temperature To date, friction mechanisms of polymer composites remian unclear due to complex composite microstructure, plastic flow behavior and effects of thermal environment. Early researchers [10, 13, 103, 104] attempt to investigate friction and wear mechanisms of engineering materials from different perspectives. However, no universal friction law has been established that can correctly describe tribological behavior of all materials. Sliding friction and wear of polymeric materials involve plastic deformation, distortion of intermolecular bonds, local material damage and fracture, interfacial material transfer and other issues related to thermal effects. Thus, sliding friction and wear of polymeric materials is a system response rather than a simple material property [44]. In the study of tribology of PTFE/PEEK composite at high temperature, friction and wear mechanisms may be more complex, critical experiments at different temperatures, as discussed in previous chapters, are needed. 9.1 Friction 9.1.1 Neat PEEK Friction of neat PEEK showed a clear transition from low to elevated temperatures (Figure 8.17). When neat PEEK polymer slides on a smooth (Ra = 0.1 µm) metallic counterfcae below its glass transition temperature, frictional force was originated primarily from adhesion between the real contact area of the PEEK and the steel counterface [104]. Bonding between the two surfaces in contact due to the adhesion was mainly weak interactions forces (e.g., hydrogen bonding and van der Waals force). The relative magnitude of the adhesion force may be expressed in terms of surface energies of 117 the polymer and the substrate. For neat PEEK, surface energy lies between 34- 38 mN/m [105]. Friction of PEEK therefore originates from continuous formations and failures of the adhesion junctions along the contact surface at temperatures below Tg. Experimental results show that very limited PEEK was transferred to the steel counterface, indicating breakage of adhesion junctions occurred at the contact interface between PEEK and steel. The mechanism of friction for neat PEEK sliding on a smooth steel counterface at low temperature was adhesive friction. Figure 9.1. SEM image of the trailing edge of PEEK slid at 200 °C. At elevated temperature, friction characteristics of neat PEEK were different from those at low temperature. Large plastic flow of neat PEEK occurred above its glass 118 transition temperature. As shown in Figure 9.1, extrusions and drawing of neat PEEK at 200 °C were clearly seen. The friction mechanism at elevated temperature was dominated by plastic deformation and flow of neat PEEK during sliding contact. Above its Tg, intermolecular sliding of PEEK polymer chains was enabled, and the polymer became easily to shear plastically. The very low modulus and yield stress at elevated temperature enabled PEEK to flow plastically under normal pressure and frictional force with large area in contact with the counterface, resulting in increased adhesion. The increased intermolecular mobility at elevated temperature resulted in a considerable transfer of PEEK to the steel counterface. The increase of adhesion at elevated temperature led to high friction of neat PEEK. 9.1.2 Neat PTFE At both low and elevated temperatures, PTFE exhibited low friction and high wear. This is mainly resulted from its unique smooth molecules and lattice crystalline structure. When PTFE slid against a steel counterface, due to mechanical shear and temperature rise, its polymer chains may have chain scission [106]. PTFE radicals generated by chain scissions then reacted with the metallic counterface and bonded chemically to the surface. With continued deposition of PTFE transfer films on the counterface, PTFE was then sliding on its own films instead of the bare steel counterface. The friction shear was shifted from the PTFE-steel interface to the PTFE-PTFE interface. The low cohesive energy of PTFE (4.19 kJ/mol [107]) and the ease of slippage between PTFE molecular chains significantly reduced the friction coefficient. Friction of PTFE was found decreased with increasing temperature from the experiment. The temperature effect on PTFE friction has been related to its viscoplastic 119 behavior [99]. The temperature dependence of PTFE friction may be modeled by a modified Arrhenius equation [51], which leads to a decrease in friction with increasing temperature. Activation energy of the sliding of PTFE was determined experimentally by Blanchet et al. [51] to be 9.2 kcal/mol and by Tanaka et al. [101] to be 7 kcal/mol. Such low activation energies indicated that van der Waals bonds were broken in the friction process of PTFE. Thus, the friction mechanism of neat PTFE is due to relative slippage between its crystallites. 9.1.3 PTFE/PEEK composite Friction of PTFE/PEEK composite at low temperature was affected by transfer films formed on the steel counterface. The overall friction measured was a resultant of friction from the film-covered area and the uncovered area. Within the transfer-film covered area, XPS analysis revealed that the transfer film consisted of both PTFE and PEEK. Onodera et al. [98] performed XPS with argon etching and molecular dynamic simulation of the transfer film for PTFE/PEEK sliding on a metallic counterface, showing that PTFE films formed on the top-most layer of the transfer films. This study provides a physical foundation for the development of current new friction theory (Eqs. 7.11 and 7.12). The friction laws, Eqs. 7.11 and 7.12, assume the film covered area has the same friction coefficient of PTFE since the top-most layer was the PTFE phase. The top layer of PTFE films was also examined with a polarized optical microscope. Figures 6.4 (b) and (c) clearly show a bright layer of PTFE on top of the transfer film for PTFE/PEEK composite and neat PTFE. Hence, when PTFE/PEEK composite is sliding on a steel counterface, transfer films formed first on the counterface with a PTFE layer on top of it. The composite then slid through the film covered area and was lubricated, due to the low 120 shear flow stress of the PTFE layer. The resulting friction of the composite was reduced and therefore was related to the amount of transfer film coverage (i.e., α). At elevated temperature, PEEK matrix of the PTFE/PEEK composite slid with large plastic deformation and flow. Friction mechanisms of PTFE/PEEK composite were changed from that observed at low temperature. Transfer films fully covered the counferface for the composites with all different compositions as shown in Figures 8.5(a) and 8.7(a). Back-scattering SEM micrograph of the composite transfer film (Figure 9.2) shows that PTFE (appears white in the image) were embedded in the film surrounded by PEEK. The PTFE at elevated temperature lubricated the composite during friction and wear. Accordingly, friction of PTFE/PEEK composite were reduced due to the lubrication effects of PTFE. Figure 9.2. SEM image (back-scattering mode) of transfer films of C15 slid on steel counterface at 200°C. 121 9.2 Wear 9.2.1 Neat PEEK Common mechanisms of sliding wear of polymers, such as thermoplastics, thermosets and elastomers are discussed by in [104]. During sliding of a polymer on a metallic counterface, surface and subsurface deformation of the polymer was caused by passage of protuberances on the counterface. The protuberance could be either asperities or debris retained in the sliding interface. Damages caused by the passing of protuberances on the polymer surface may be adhesive and/or abrasive. These two wear modes are not mutually exclusive and usually coexist depending on materials of the sliding pair and sliding conditions [39]. In general, characteristics of abrasive wear are scratching marks on the polymer surface and those of adhesive wear are scale-like wear particles. In the case of neat PEEK sliding on steel counterface at low temperature, both scratching marks and scale-like wear debris were seen (Figure 8.4), suggesting the wear mechanism of neat PEEK at low temperature may be a combination of adhesive and abrasive wear. At elevated temperature, the specific wear rate of neat PEEK reduced pronouncedly. Associated with the low wear, transfer films were seen on the counterface (Figures 8.5 (a) and (c)). In addition, polymer drawing at the trailing edge of the PEEK specimen indicated large plastic flow. As a result, PEEK was transferred to the steel counterface during sliding and stayed. Instead of fracture and forming wear debris during the sliding of PEEK at low temperature, it exhibited large deformation and plastic flow at elevated temperature. The net material loss was minimal and resulted in a low wear rate. 122 The mechanisms responsible for its low wear was mainly by transfer and plastic flow of PEEK. 9.2.2 Neat PTFE Neat PTFE had high wear rate at both low and elevated temperatures. The wear mechanism of PTFE is related to its unique lattice crystalline structure and smooth polymer chain profile. Figure 8.8, long films can be seen generated at the trailing edge during sliding of PTFE on a steel surface. The film formation and polymer drawing were attributed to the destruction of lattice structure by slippage of crystalline slices. The transferred PTFE on the counterface was then get scraped by the PTFE pins in the next revolution of rotation and accumulated at the leading edge of the pin (Figures 8.7 (b) and (e)). This transfer-scrape process repeated itself during the sliding of PTFE and eventually led to high wear. 9.2.3 PTFE/PEEK Composite Wear of PTFE/PEEK composite was reduced because of the presence of transfer films. At low temperature, transfer films on the sliding counterface firstly prevented direct contact of the steel counterface from the PTFE/PEEK composite sliding pin. Though transfer films were thin and only partially covered the steel counterface, abrasion of polymer composite surface is mitigated. Less scratching marks were seen on the composite worn surface (Figure 8.6 (d)) than that on neat PEEK worn surface (Figure 8.4(d)). Also, the aforementioned PTFE layer on top of the transfer film reduced adhesion and friction shear on the composite. The adhesive wear of the composite was not as significant as that of the neat PEEK, which resulted in a lowered wear rate. 123 At elevated temperature, wear rates of PTFE/PEEK composite were similar as shown in Figure 8.18. This may be attributed to plastic flows of both PEEK and PTFE. Once the PEEK and the PTFE in the composite started to deform plastically and were transferred to the counterface during sliding, the transferred materials tended to stay on the counterface. Loose wear particles and debris were rarely seen during the sliding. Therefore, net wear loss of the material was low and such a low wear rate was observed. 124 Chapter 10 Conclusions A combined experimental and theoretical study on tribological behavior of PTFE/PEEK composite at elevated temperature has been carried out. Based on the results, the following conclusions may be drawn: (1) Friction coefficients of PTFE/PEEK composite at low temperature decreased with increasing PTFE volume fraction. This may attribute to transfer film lubrication. The transfer film coverage ratio was found to relate to the volume fraction of the PTFE phase by an error function. (2) Below glass transition temperature of PEEK (152 °C), friction coefficients of PTFE/PEEK composites decreased with increasing temperature due to lubrication of PTFE. Above glass transition temperature of PEEK, friction coefficients of the composite increased with temperature due to plastic deformation and flow of the PEEK matrix and the PTFE phase. (3) Specific wear rates of PTFE/PEEK composite at low temperature decreased with PTFE volume fraction due to increased transfer film coverage, which prevented direct abrasion of the composite surface and mitigated adhesion between polymer composite and the counterface. (4) Specific wear rates of PTFE/PEEK composites at elevated temperature above glass transition of PEEK were low and their values were similar. At elevated temperature, PEEK started to flow plastically and stayed on the counterface instead of becoming wear debris and expelled. 125 (5) Friction and wear of PTFE/PEEK composite are inherently related. A power law relationship between the two is established at both low and elevated temperatures. (6) At low temperature, a new friction theory is developed with the aid of solid film lubrication and the rule of the mixtures for the PTFE/PEEK composite. With the aid of the transfer film cover age ratio, the theory is shown to predict friction coefficients of PTFE/PEEK composite at low temperature. (7) Friction and wear mechanisms of PTFE/PEEK composite are fundamentally different below and above the PEEK glass transition temperature. A new mechanism-based friction model is introduced for developing PTFE/PEEK composite friction and wear laws at elevated temperature. Theoretical predictions show good agreement with elevated temperature experimental results. This model also provides good correlation with test data at low temperature. Thus, this model is more general and applicable to all temperatures. (8) Wear of PTFE/PEEK composite at both low and elevated temperatures are modeled with a power law relationship. The new wear theory is used to predict specific wear rates of the PTFE/PEEK composite at both low and elevated temperatures. The combined experimental and theoretical investigation carried out in this study provides a new perspective to investigate tribological behavior of PTFE/PEEK composite 126 at both low and elevated temperatures. Merits and innovative contributions of this investigation to the PTFE/PEEK composite tribological behavior include the following: (1) Extending the friction and wear coefficient power-law relationship observed in metal and non-metal to polymer composite materials with the use of specific wear rate to characterize polymeric material wear behavior. (2) Develop PTFE/PEEK composite friction and wear models to enable the use of only constituent properties to predict friction and wear of PTFE/PEEK composite. (3) Identify differences in friction and wear behavior of PTFE/PEEK composite at low and elevated temperatures. (4) Provision of new research direction that includes both PTFE morphology and composite microstructure in friction and wear predictions. Critical issues that have not been fully addressed in this study require further investigations including but not limited to: (1) Physical interpretation and clear understanding of the parameters used in friction and wear modeling, i.e., β and ξ. (2) Effect of viscoplastic behavior of neat PEEK, neat PTFE and PTFE/PEEK composites on friction and wear at elevated temperature. (3) Rigorous quantitative analysis and evaluation of the relationship between PTFE crystalline morphology and composite microstructure, such as crystal size, orientation and degrees of crystallinity, and elevated temperature friction and wear of PTFE/PEEK composites at both low and elevated temperatures. 127 References 1. Jost, H.P.: “The introduction of a new technology, report from the committee on tribology.” Her Majesty’s Station. Off. London. (1966) 2. Lee, P.M., Carpick, R.: “Tribological opportunities for enhancing America’s energy efficiency.” A Rep. to Adv. Res. Proj. Agency-Energy US Dep. Energy. 14, (2017) 3. Holmberg, K., Erdemir, A.: “Influence of tribology on global energy consumption, costs and emissions.” Friction. 5, 263–284 (2017). doi:10.1007/s40544-017-01835 4. Layard, A.H.: Discoveries among the Ruins of Nineveh and Babylon. Harper & Bros. (1853) 5. Dowson, D.: History of tribology. Addison-Wesley Longman Limited (1979) 6. Amontons, G.: “De la Résistance Causée dans les Machines.” Mémoires l’Académie R. A. 275–282 (1699) 7. Coulomb, C.A.: “Théorie des machines simples, en ayant égard au frottement de leurs parties et à la roideur des cordages.” Mem. Math. Phys. X, 161–342 (1785) 8. Huber, M.T.: “Zur Theorie der Berührung fester elastischer Körper.” Ann. Phys. 319, 153–163 (1904). doi:10.1002/andp.19043190611 9. Bowden, F.P., Tabor, D.: The friction and lubrication of solids. 2nd corrected ed. Oxford Univ. Press. Oxford. (1986) 10. Suh, N.P.: Tribophysics. Pretice-Hall, Englewood Cliffs, NJ, 1986. (1986) 128 11. Wilson, W.R.D., Sheu, S.: “Real area of contact and boundary friction in metal forming.” Int. J. Mech. Sci. 30, 475–489 (1988). doi:10.1016/00207403(88)90002-1 12. Komvopoulos, K., Suh, N.P., Saka, N.: “Wear of boundary-lubricated metal surfaces.” Wear. 107, 107–132 (1986). doi:10.1016/0043-1648(86)90022-0 13. Rabinowicz, E.: Friction and Wear of Materials. John Wiley & Sons, New York, NY (1965) 14. Pei, X., Friedrich, K.: “Sliding wear properties of PEEK, PBI and PPP.” Wear. 274, 452–455 (2012) 15. Yamaguchi, Y.: Tribology of plastic materials: their characteristics and applications to sliding components. Elsevier Science Publishers B.V. Amsterdam the Netherland (1990) 16. Stuart, B.H.: “Tribological studies of poly (ether ether ketone) blends.” Tribol. Int. 31, 647–651 (1998) 17. Laux, K.A., Schwartz, C.J.: “Influence of linear reciprocating and multi-directional sliding on PEEK wear performance and transfer film formation.” Wear. 301, 727– 734 (2013). doi:10.1016/J.WEAR.2012.12.004 18. Wieleba, W.: “The statistical correlation of the coefficient of friction and wear rate of PTFE composites with steel counterface roughness and hardness.” Wear. 252, 719–729 (2002). doi:10.1016/S0043-1648(02)00029-7 19. Burris, D.L., Sawyer, W.G.: “A low friction and ultra low wear rate PEEK/PTFE 129 composite.” Wear. 261, 410–418 (2006). doi:10.1016/J.WEAR.2005.12.016 20. Zhao, X., Hamilton, M., Sawyer, W.G., Perry, S.S.: “Thermally Activated Friction.” Tribol. Lett. 27, 113–117 (2007). doi:10.1007/s11249-007-9220-2 21. Kurdi, A., Chang, L., Kurdi, A., Chang, L.: “Recent Advances in High Performance Polymers—Tribological Aspects.” Lubricants. 7, 2 (2018). doi:10.3390/lubricants7010002 22. Ovaert, T.C., Ramachandra, S.: “The effect of controlled counterface topography on polymer transfer and wear.” Int. J. Mach. Tools Manuf. 35, 311–316 (1995). doi:10.1016/0890-6955(94)P2388-V 23. Bahadur, S., Sunkara, C.: “Effect of transfer film structure, composition and bonding on the tribological behavior of polyphenylene sulfide filled with nano particles of TiO2, ZnO, CuO and SiC.” Wear. 258, 1411–1421 (2005). doi:10.1016/J.WEAR.2004.08.009 24. Bahadur, S.: “The development of transfer layers and their role in polymer tribology.” Wear. 245, 92–99 (2000). doi:10.1016/S0043-1648(00)00469-5 25. Ye, J., Haidar, D., Burris, D.: “Polymeric solid lubricant transfer films: Relating quality to wear performance.” In: Self-Lubricating Composites. pp. 155–180. Springer Berlin Heidelberg, Berlin, Heidelberg (2018) 26. Qu, S., Lo, K.H., Wang, S.-S.: “Wear of PTFE /PEEK Composite: Theory and Experiments.” In: Abstracts of Papers of the 2018 STLE Tribology Frontiers Conference, Chicago IL, Oct. 28-31 (2018) 130 27. Burris, D.L., Sawyer, W.G.: “Tribological behavior of PEEK components with compositionally graded PEEK/PTFE surfaces.” Wear. 262, 220–224 (2007). doi:10.1016/J.WEAR.2006.03.045 28. Lu, Z.P., Friedrich, K.: “On sliding friction and wear of PEEK and its composites.” Wear. 181–183, 624–631 (1995). doi:10.1016/0043-1648(95)90178-7 29. Hufenbach, W., Kunze, K., Bijwe, J.: “Sliding wear behaviour of PEEK-PTFE blends.” J. Synth. Lubr. 20, 227–240 (2003). doi:10.1002/jsl.3000200305 30. Brazel, C.S., Rosen, S.L.: Fundamental principles of polymeric materials. John Wiley & Sons, Hoboken, NJ (2012) 31. Dowson, P., Walker, M.S., Watson, A.P.: “Development of abradable and rubtolerant seal materials for application in centrifugal compressors and steam turbines.” Seal. Technol. 2004, 5–10 (2004). doi:10.1016/S1350-4789(04)00451-9 32. Friedrich, K., Lu, Z., Häger, A.M.: “Overview on polymer composites for friction and wear application.” Theor. Appl. Fract. Mech. 19, 1–11 (1993). doi:10.1016/0167-8442(93)90029-B 33. Duan, Y., Cong, P., Liu, X., Li, T.: “Friction and wear of polyphenylene sulfide (PPS), polyethersulfone (PES) and polysulfone (PSU) under different cooling conditions.” J. Macromol. Sci. Part B Phys. 48, 604–616 (2009). doi:10.1080/00222340902837899 34. Mens, J.W.M., de Gee, A.W.J.: “Friction and wear behaviour of 18 polymers in contact with steel in environments of air and water.” Wear. 149, 255–268 (1991). 131 doi:10.1016/0043-1648(91)90378-8 35. Saikko, V.: “Effect of Contact Area on the Wear and Friction of UHMWPE in Circular Translation Pin-on-Disk Tests.” J. Tribol. 139, 061606 (2017). doi:10.1115/1.4036448 36. Sidebottom, M.A., Pitenis, A.A., Junk, C.P., Kasprzak, D.J., Blackman, G.S., Burch, H.E., Harris, K.L., Sawyer, W.G., Krick, B.A.: “Ultralow wear Perfluoroalkoxy (PFA) and alumina composites.” Wear. 362–363, 179–185 (2016). doi:10.1016/j.wear.2016.06.003 37. Yang, Z., Dong, B., Huang, Y., Liu, L., Yan, F.-Y., Li, H.-L.: “Enhanced wear resistance and micro-hardness of polystyrene nanocomposites by carbon nanotubes.” Mater. Chem. Phys. 94, 109–113 (2005). doi:10.1016/J.MATCHEMPHYS.2005.04.029 38. Bhimaraj, P., Burris, D.L., Action, J., Sawyer, W.G., Toney, C.G., Siegel, R.W., Schadler, L.S.: “Effect of matrix morphology on the wear and friction behavior of alumina nanoparticle/poly(ethylene) terephthalate composites.” Wear. 258, 1437– 1443 (2005). doi:10.1016/J.WEAR.2004.09.077 39. Friedrich, K., Reinicke, P.: “Friction and wear of polymer-based composites.” Mech. Compos. Mater. 34, 503–514 (1998). doi:10.1007/BF02254659 40. Shi, G., Zhang, M.Q., Rong, M.Z., Wetzel, B., Friedrich, K.: “Sliding wear behavior of epoxy containing nano-Al2O3 particles with different pretreatments.” Wear. 256, 1072–1081 (2004). doi:10.1016/S0043-1648(03)00533-7 132 41. Sawyer, W.G., Freudenberg, K.D., Bhimaraj, P., Schadler, L.S.: “A study on the friction and wear behavior of PTFE filled with alumina nanoparticles.” Wear. 254, 573–580 (2003). doi:10.1016/S0043-1648(03)00252-7 42. Zhang, X.-R., Pei, X.-Q., Wang, Q.-H.: “Friction and wear studies of polyimide composites filled with short carbon fibers and graphite and micro SiO2.” Mater. Des. 30, 4414–4420 (2009). doi:10.1016/J.MATDES.2009.04.002 43. Cao, W., Gong, J., Yang, D., Gao, G., Wang, H., Ren, J., Chen, S.: “Tribological behavior and energy dissipation characteristics of nano-Al2O3-reinforced PTFEPPS composites in sliding system.” J. Cent. South Univ. 24, 2001–2009 (2017). doi:10.1007/s11771-017-3609-3 44. Kurdi, A., Kan, W.H., Chang, L.: “Tribological behaviour of high performance polymers and polymer composites at elevated temperature.” Tribol. Int. 130, 94– 105 (2019). doi:10.1016/J.TRIBOINT.2018.09.010 45. Lancaster, J.K.: “Polymer-based bearing materials: The role of fillers and fibre reinforcement.” Tribology. 5, 249–255 (1972). doi:10.1016/0041-2678(72)90103-0 46. Burris, D.L., Santos, K., Lewis, S.L., Liu, X., Perry, S.S., Blanchet, T.A., Schadler, L.S., Sawyer, W.G.: “Polytetrafluoroethylene matrix nanocomposites for tribological applications.” Tribol. Interface Eng. Ser. 55, 403–438 (2008). doi:10.1016/S1572-3364(08)55017-8 47. Stachowiak, G. W., Batchelor, A.W.: Engineering tribology. ButterworthHeinemann, Burlington, MA (2013) 133 48. Wang, Q., Chung, Y. Eds.: Encyclopedia of tribology. Springer (2013) 49. Pooley, C.M., Tabor, D.: “Friction and Molecular Structure: The Behaviour of Some Thermoplastics.” Proc. R. Soc. A Math. Phys. Eng. Sci. 329, 251–274 (1972). doi:10.1098/rspa.1972.0112 50. Pooley, C.M., Tabor, D.: “Transfer of PTFE and Related Polymers in a Sliding Experiment.” Nat. Phys. Sci. 237, 88–90 (1972). doi:10.1038/physci237088a0 51. Blanchet, T.A., Kennedy, F.E.: “Sliding wear mechanism of polytetrafluoroethylene (PTFE) and PTFE composites.” Wear. 153, 229–243 (1992). doi:10.1016/0043-1648(92)90271-9 52. Makinson, K. R., Tabor, D.: “Friction and transfer of polytetrafluoroethylene.” Nature. 201, 1172, 464-466 (1964). doi:10.1038/201464a0 53. Shi, Y.J., Feng, X., Wang, H.Y., Liu, C., Lu, X.H.: “Effects of filler crystal structure and shape on the tribological properties of PTFE composites.” Tribol. Int. 40, 1195–1203 (2007). doi:10.1016/J.TRIBOINT.2006.12.006 54. McElwain, S.E., Blanchet, T.A., Schadler, L.S., Sawyer, W.G.: “Effect of Particle Size on the Wear Resistance of Alumina-Filled PTFE Micro- and Nanocomposites.” Tribol. Trans. 51, 247–253 (2008). doi:10.1080/10402000701730494 55. Lancaster, J.K.: “The effect of carbon fibre reinforcement on the friction and wear of polymers.” J. Phys. D. Appl. Phys. 1, 303 (1968). doi:10.1088/00223727/1/5/303 134 56. Tanaka, K., Kawakami, S.: “Effect of various fillers on the friction and wear of polytetrafluoroethylene-based composites.” Wear. 79, 221–234 (1982). doi:10.1016/0043-1648(82)90170-3 57. Khedkar, J., Negulescu, I., Meletis, E.I.: “Sliding wear behavior of PTFE composites.” Wear. 252, 361–369 (2002). doi:10.1016/S0043-1648(01)00859-6 58. Li, F., Hu, K., Li, J., Zhao, B.: “The friction and wear characteristics of nanometer ZnO filled polytetrafluoroethylene.” Wear. 249, 877–882 (2001). doi:10.1016/S0043-1648(01)00816-X 59. Kandanur, S.S., Schrameyer, M.A., Jung, K.F., Makowiec, M.E., Bhargava, S., Blanchet, T.A.: “Effect of Activated Carbon and Various Other Nanoparticle Fillers on PTFE Wear.” Tribol. Trans. 57, 821–830 (2014). doi:10.1080/10402004.2014.916374 60. Briscoe, B.J., Lin Heng Yao, Stolarski, T.A.: “The friction and wear of poly(tetrafluoroethylene)-poly (etheretherketone) composites: An initial appraisal of the optimum composition.” Wear. 108, 357–374 (1986). doi:10.1016/00431648(86)90013-X 61. Bijwe, J., Sen, S., Ghosh, A.: “Influence of PTFE content in PEEK–PTFE blends on mechanical properties and tribo-performance in various wear modes.” Wear. 258, 1536–1542 (2005). doi:10.1016/J.WEAR.2004.10.008 62. Lal, B., Alam, S., Mathur, G.N.: “Tribo-investigation on PTFE lubricated PEEK in harsh operating conditions.” Tribol. Lett. 25, 71–77 (2006). doi:10.1007/s11249006-9158-9 135 63. Vail, J.R., Krick, B.A., Marchman, K.R., Sawyer, W.G.: “Polytetrafluoroethylene (PTFE) fiber reinforced polyetheretherketone (PEEK) composites.” Wear. 270, 737–741 (2011). doi:10.1016/j.wear.2010.12.003 64. Qu, S., Penaranda, J., Wang, S.S. “Tribological behavior of PTFE/PEEK composite.” In: Proceedings of the American Society for Composites - 31st Technical Conference ,Davidson, B.D., Ratclif, J.G. Czabaj Eds, Williamsburg, VA, Sept. 19- 21. (2016) 65. Qu, S., Peneranda, J., Wang, S.S.: “Elevated-Temperature Wear and Friction of PTFE/PEEK composite.” In: 32nd Technical Conference of the American Society for Composites, Purdue University, Lafayette, IN, Oct. 23-25. (2017) 66. Ye, J., Burris, D., Xie, T., Ye, J., Burris, D.L., Xie, T.: “A Review of Transfer Films and Their Role in Ultra-Low-Wear Sliding of Polymers.” Lubricants. 4, 4 (2016). doi:10.3390/lubricants4010004 67. Wang, Y., Yan, F.: “Tribological properties of transfer films of PTFE-based composites.” Wear. 261, 1359–1366 (2006). doi:10.1016/j.wear.2006.03.050 68. Makinson, K.R., Tabor, D.: “The friction and transfer of polytetrafluoroethylene.” Proc. R. Soc. London. Ser. A. Math. Phys. Sci. 281, 49–61 (1964). doi:10.1098/rspa.1964.0168 69. Biswas, S.K., Vijayan, K.: “Changes to near-surface region of PTFE during dry sliding against steel.” J. Mater. Sci. 23, 1877–1885 (1988). doi:10.1007/BF01115734 136 70. Vande Voort, J., Bahadur, S.: “The growth and bonding of transfer film and the role of CuS and PTFE in the tribological behavior of PEEK.” Wear. 181–183, 212–221 (1995). doi:10.1016/0043-1648(95)90026-8 71. Bayer, R.G. Ed.: STP701-EB Wear Tests for Plastics: Selection and Use. ASTM International, West Conshohocken, PA (1979) 72. Benzing, R.J., Hopkins, V., Villforth, F., Petronio, M.: Friction and Wear DevicesA Survey. US Air Force Materials Laboratory Technical Memorandum MAN-6616 (1966) 73. ASTM G137-97(2017) Standard Test Method for Ranking Resistance of Plastic Materials to Sliding Wear Using a Block-On-Ring Configuration, ASTM International, West Conshohocken, PA (2003) 74. ASTM G99-17 Standard Test Method for Wear Testing with a Pin-on-Disk Apparatus, ASTM International, West Conshohocken (2011) 75. ASTM G133-05(2016) Standard Test Method for Linearly Reciprocating Ball-onFlat Sliding Wear, ASTM International, West Conshohocken (2011) 76. ASTM D3702-94(2014) Standard Test Method for Wear Rate and Coefficient of Friction of Materials in Self- Lubricated Rubbing Contact Using a Thrust Washer Testing, ASTM International, West Conshohocken (1999) 77. Khruschov, M.M.: “Principles of abrasive wear.” Wear. 28, 69–88 (1974). doi:10.1016/0043-1648(74)90102-1 78. Simm, W., Freti, S.: “Abrasive wear of multiphase materials.” Wear. 129, 105–121 137 (1989). doi:10.1016/0043-1648(89)90283-4 79. Liu, G., Wang, S.S.: “Multi-Axial Yielding, Plastic Flow and Failure Strength of PTFE/PEEK Composites.” In: Proceedings of 29th Technical Conference, American Society for Composites. 109 Kim. H., Whisler, D., Chen, Z-M., Bisagni, C. Kawai, M. Kruger, R. Eds., La Jolla, CA. Sept. 8-10 (2014) 80. Ruland, W.: “X-ray determination of crystallinity and diffuse disorder scattering.” Acta Crystallogr. 14, 1180–1185 (1961). doi:10.1107/S0365110X61003429 81. Blundell, D.J., Osborn, B.N.: “The morphology of poly(aryl-ether-ether-ketone).” Polymer (Guildf). 24, 953–958 (1983). doi:10.1016/0032-3861(83)90144-1 82. Ryland, A.L.: X-ray diffraction, Polychemicals Department, E.I. Du Pont de Nemours & Co, Inc. Willmington, DE (1958) 83. Hiemenz, P.C., Lodge, T.P.: Polymer chemistry. CRC press (2007) 84. ASTM D785-08-Standard Test Method for Rockwell Hardness of Plastics and Electrical Insulating Materials. ASTM International, West Conshohocken, PA (2015) doi:10.1520/D0785-08.2 85. ASTM D7028 - 07 (2015) Standard Test Method for Glass Transition Temperature (DMA Tg) of Polymer Matrix Composites by Dynamic Mechanical Analysis (DMA),” ASTM International, West Conshohocken, PA (2015) 86. Sundarara, S.: “Material Wear Evaluation using Temperature Controlled Wear Testing,” SAE International Journal of Materials and Manufacturing 6, 339-348 (2013) doi: 10.4271/2013-01-1218 138 87. Sundararaman, S.: “Improved Test Method for Tribological Evaluation of High Performance Plastics.” In SAE International Paper#2019-01-0183, WCX SAE World Congress Experience, Detroit, MI (2019) 88. Haidar, D.R., Alam, K.I., Burris, D.L.: “Tribological Insensitivity of an UltralowWear Poly(etheretherketone)-Polytetrafluoroethylene Polymer Blend to Changes in Environmental Moisture.” J. Phys. Chem. C. 122, 5518–5524 (2018). doi:10.1021/acs.jpcc.7b12487 89. Bahadur, S., Gong, D.: “The action of fillers in the modification of the tribological behavior of polymers.” Wear. 158, 41–59 (1992). doi:10.1016/00431648(92)90029-8 90. Li, X., Gao, Y., Xing, J., Wang, Y., Fang, L.: “Wear reduction mechanism of graphite and MoS2 in epoxy composites.” Wear. 257, 279–283 (2004). doi:10.1016/J.WEAR.2003.12.012 91. Bahadur, S., Tabor, D.: “The wear of filled polytetrafluoroethylene.” Wear. 98, 1– 13 (1984). doi:10.1016/0043-1648(84)90213-8 92. Schwartz, C.J., Bahadur, S.: “Studies on the tribological behavior and transfer film–counterface bond strength for polyphenylene sulfide filled with nanoscale alumina particles.” Wear. 237, 261–273 (2000). doi:10.1016/S00431648(99)00345-2 93. Bhimaraj, P., Burris, D., Sawyer, W.G., Toney, C.G., Siegel, R.W., Schadler, L.S.: “Tribological investigation of the effects of particle size, loading and crystallinity on poly(ethylene) terephthalate nanocomposites.” Wear. 264, 632–637 (2008). 139 doi:10.1016/J.WEAR.2007.05.009 94. Wang, Q., Xue, Q., Liu, H., Shen, W., Xu, J.: “The effect of particle size of nanometer ZrO2 on the tribological behaviour of PEEK.” Wear. 198, 216–219 (1996). doi:10.1016/0043-1648(96)07201-8 95. Wang, Q.-H., Xu, J., Shen, W., Xue, Q.: “The effect of nanometer SiC filler on the tribological behavior of PEEK.” Wear. 209, 316–321 (1997). doi:10.1016/S00431648(97)00015-X 96. Wang, Q., Xue, Q., Shen, W.: “The friction and wear properties of nanometre SiO2 filled polyetheretherketone.” Tribol. Int. 30, 193–197 (1997). doi:10.1016/S0301-679X(96)00042-4 97. Kalin, M., Zalaznik, M., Novak, S.: “Wear and friction behaviour of poly-etherether-ketone (PEEK) filled with graphene, WS2 and CNT nanoparticles.” Wear. 332–333, 855–862 (2015). doi:10.1016/J.WEAR.2014.12.036 98. Onodera, T., Nunoshige, J., Kawasaki, K., Adachi, K., Kurihara, K., Kubo, M.: “Structure and Function of Transfer Film Formed from PTFE/PEEK Polymer Blend.” J. Phys. Chem. C. 121, 14589–14596 (2017). doi:10.1021/acs.jpcc.7b02860 99. McCook, N.L., Burris, D.L., Dickrell, P.L., Sawyer, W.G.: “Cryogenic Friction Behavior of PTFE based Solid Lubricant Composites.” Tribol. Lett. 20, 109–113 (2005). doi:10.1007/s11249-005-8300-4 100. Friedrich, K.: Wear models for multiphase materials and synergistic effects in 140 polymer hybrid composite, Chapt. 6 Advances in Composite Tribology, Friedrich, K. ed., 209-276, Elsevier Science Publishers B.V. Amsterdam, the Netherland (1993) 101. Tanaka, K., Uchiyama, Y., Toyooka, S.: “The mechanism of wear of polytetrafluoroethylene.” Wear. 23, 153–172 (1973). doi:10.1016/00431648(73)90081-1 102. Babuska, T.F., Pitenis, A.A., Jones, M.R., Nation, B.L., Sawyer, W.G., Argibay, N.: “Temperature-Dependent Friction and Wear Behavior of PTFE and MoS2.” Tribol. Lett. 63, 15 (2016). doi:10.1007/s11249-016-0702-y 103. Bartenev, G.M., LavrentΚΉev, V.V.: Friction and wear of polymers. Lee, L.-H., Ludema, K.C. Eds., Elsevier Scientific Pub. Co. Amsterdam, the Netherland (1981) 104. Hutchings, I.M., Shipway, P.: Tribology : friction and wear of engineering materials. 2nd ed., Butterworth-Heinemann, Kidlington, UK (2017) 105. Victrex PEEK Film technology: Surface Treatment and Adhesion of PEEK Film. downloaded July 13th 2019 https://www.victrex.com/~/media/literature/en/ 106. Biswas, S.K., Vijayan, K.: “Friction and wear of PTFE — a review.” Wear. 158, 193–211 (1992). doi:10.1016/0043-1648(92)90039-B 107. Giltrow, J.P.: “A relationship between abrasive wear and the cohesive energy of materials.” Wear. 15, 71–78 (1970). doi:10.1016/0043-1648(70)90187-0 141 Appendix A Numerical Simulation of Randomly Distributed Spherical Particle Filled Composite In this study, the cross-sectional area fraction of particles in a composite was investigated by a numerical simulation method. For a given randomly distributed spherical particles reinforced composite, the correlation between volumetric fraction of the filler and the area fraction of particles on a cross section was studied. Models of composite with different filler volume fraction was created and the cross-sectional area fractions were calculated and compared with the volume fraction. The results showed that the crosssectional area fraction is approximately equal to the volume fraction if the particle size is small and total number of particles is large. To construct the model, following constrains were implemented. 1. Radii of particles are normally distributed with a mean and a standard deviation. 2. Locations of particle are randomly distributed such as the three coordinates (x, y, and z) are totally random. 3. Particles are mutually exclusive. Steps to construct the composite model were illustrated in the flowchart below in Figure A.1. 142 Figure A.1. Flow chart of the particle reinforced composite model. Once the model is constructed, 100 slices were taken along z-axis. The slicing process is shown in Figure A.2 and an example of a slice of cross section is shown in Figure A.3. 143 Figure A.2. Illustration of slicing the cross section. Figure A.3. A cross section of 10% by volume particle filled composite. 144 For each slice of cross section, the area fraction is calculated and compared with the composite volume fraction. The particle size effect was also investigated. Particle mean radii ranging from 0.5 to 4 were tested for various volume fractions up to 20%. A 10% volume fraction example is shown in Figure A.4. Figure A.4. Distribution of the area to volume fraction ratio of a 10% composite for different particle mean radii. As shown in in the figure above, with shirking the particle size and increasing particle number, the distribution of the area to volume fraction ratio coverages to 1 gradually. The area to volume fraction ratio of other volume fraction composites in this simulation study exhibited the same converging behavior. For a particle radius to matrix length ratio of 1/100, the mean of area to volume fraction ratio was 0.9959 with a standard deviation of 0.0453. For PTFE/PEEK composites, the average PTFE particle radius was 50 microns and the cubic pin was 6.35 mm each side. That yielded a particle radius to matrix length 145 ratio of 1/127, which is close to the value in the simulation. Therefore, according to the numerical simulation, it is safe to claim that the cross-sectional area fraction is approximately equal to the volume fraction of the reinforcing particle. 146 Appendix B Friction Coefficients by the Rule of Mixtures To include the effect of transfer films on the friction behavior of PTFE/PEEK composites, the contact area (test pin cross-section area) between the composite test pin and steel counterface is divided into two distinctly different areas – one area with and one area without transfer films coverage. Assuming the shear stress (π) across the contact area (π΄) is uniformly distributed, the total shear force (πΉπ ) acting on the contact area is equal to ππ΄. The shear force (πΉππΉ ) acting on the contact area covered by transfer films is given by πΉππΉ = ππΌπ΄ , (π΅. 1) where πΌ is the transfer films area coverage ratio (ACR). Similarly, in the contact area not covered by transfer films, the shear forces carried by the PTFE particles (πΉππ ) and the PEEK matrix (πΉππΎ ) are given by πΉππ = π(1 − πΌ)ππ π΄ and πΉππΎ = π(1 − πΌ)(1 − ππ )π΄ , (π΅. 2) where ππ is the PTFE volume fraction in the composite. Now the total normal load (πΏπ ) acting in the contact area is given by πΏπ = πΏππΉ + πΏππ + πΏππΎ , (π΅. 3) where πΏππΉ is the normal load on the contact area covered by transfer films, and πΏππ and πΏππΎ are the normal loads carried by the PTFE particles and the PEEK matrix in the contact 147 area not covered by transfer films. Using the relationship between friction coefficient (π), shearing force (πΉ) and normal load (πΏ), Eq. (18) can be written as πΉπ πΉππΉ πΉππ πΉππΎ = + + . ππ π ππΉ πππ πππΎ (π΅. 4) Substituting Eqs. (16) and (17) into Eq. (19), we have the following expression for the friction coefficient (ππ ) of PTFE/PEEK composites (1 − πΌ)ππ (1 − πΌ)(1 − ππ ) 1 πΌ = + + ππ π ππΉ πππ πππΎ (π΅. 5) or ππ = 1 (1 − πΌ)ππ (1 − πΌ)(1 − ππ ) πΌ + + π ππΉ πππ πππΎ . (π΅. 6) Equation (21) is the inverse rule of mixture (IROM) for the friction coefficient of PTFE/PEEK composites. If a uniform pressure distribution on the contact area is assumed instead of a uniform shear stress distribution for the determination of composite friction coefficient, the following linear rule of mixture (LROM) is obtained for ππ , i.e., ππ = πΌπ ππΉ + (1 − πΌ)ππ πππ + (1 − πΌ)(1 − ππ )πππΎ . 148 (π΅. 7)