© 2016 kathryn leigh harris - university of...

107
TRIBOCHEMICAL INTERACTIONS OF A PTFE/ALPHA ALUMINA COMPOSITE AT THE SLIDING INTERFACE: A MECHANISM FOR ULTRA LOW WEAR By KATHRYN LEIGH HARRIS A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2016

Upload: others

Post on 10-Jun-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

TRIBOCHEMICAL INTERACTIONS OF A PTFE/ALPHA ALUMINA COMPOSITE AT

THE SLIDING INTERFACE: A MECHANISM FOR ULTRA LOW WEAR

By

KATHRYN LEIGH HARRIS

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL

OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2016

Page 2: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

© 2016 Kathryn Leigh Harris

Page 3: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

To Infinity and Beyond

Page 4: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

4

ACKNOWLEDGMENTS

I would like to thank our collaborators at DuPont for their support of this research. In

particular, Gregory S. Blackman and Christopher P. Junk have been invaluable for their generous

contribution of time and effort over the last few years. I would also like to thank the PTFE

enthusiasts whose efforts in the area preceded, coincided with and contributed to the work

presented here, including Professors David Burris and Brandon Krick, and Dr. Angela Pitenis. I

am forever grateful to all past and present members of the UF Tribology Lab for their dedication

in constructing a truly one of a kind experience in teamwork and friendship.

I thank my advisor, Dr. Gregory Sawyer for not only his academic guidance, but also for

the friendship, strength and personal support I have experienced under his leadership that have

known no equal, and are unlikely to.

Page 5: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

5

TABLE OF CONTENTS

page

ACKNOWLEDGMENTS ...............................................................................................................4

LIST OF TABLES ...........................................................................................................................7

LIST OF FIGURES .........................................................................................................................8

ABSTRACT ...................................................................................................................................11

CHAPTER

1 INTRODUCTION ..................................................................................................................13

2 BACKGROUND ....................................................................................................................15

The History of Polytetrafluoroethylene ..................................................................................15 The Wear of Neat PTFE .........................................................................................................17

Friction and Wear of PTFE Composites .................................................................................19 Tribofilms and Ultra-Low Wear .............................................................................................21 The Effects of Surface Roughness ..........................................................................................23

The Effects of α-Alumina Particle Morphology .....................................................................27 Tribochemistry ........................................................................................................................31

3 METHODS AND EXPERIMENTATION .............................................................................45

Materials and Sample Preparation ..........................................................................................45

Tribometer and Wear Test Design ..........................................................................................46 Wear Rate Calculation, Friction Measurements and Uncertainty ..........................................46

Stylus Profilometry .................................................................................................................48 Small Molecules Experiments ................................................................................................49

X-Ray Photoelectron Spectroscopy ........................................................................................49 Infrared Spectroscopy .............................................................................................................50 Etched PTFE Tests .................................................................................................................51

4 RESULTS AND DISCUSSION .............................................................................................54

Friction and Wear ...................................................................................................................54

Stylus Profilometry .................................................................................................................54 X-Ray Photoelectron Spectroscopy ........................................................................................54

Infrared Spectroscopy .............................................................................................................56

5 CONCLUSIONS ....................................................................................................................72

APPENDIX: FUTURE CONSIDERATIONS: COUNTERFACE EFFECTS..............................75

Page 6: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

6

Wear and Friction Experiments on Additional Countersurfaces ............................................75

Infrared Spectroscopy and X-Ray Analysis of Transfer and Running Films .........................78 Surface and Sub-Surface Evolution of the Aluminum Countersamples ................................80

LIST OF REFERENCES ...............................................................................................................99

BIOGRAPHICAL SKETCH .......................................................................................................107

Page 7: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

7

LIST OF TABLES

Table page

2-1 Tensile and compressive yield strengths of unfilled PTFE as reported by *Rae109 and

**DuPont110. ......................................................................................................................34

2-2 A summary of the particle size results, which vary by method, and the wear rate of

the PTFE composite containing them ................................................................................40

Page 8: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

8

LIST OF FIGURES

Figure page

2-1 The polymerization reaction of PTFE................................................................................33

2-2 The PTFE molecule is helical in structure .........................................................................33

2-3 PTFE crystallizes to form lamellae rather than spherulites ...............................................34

2-4 A suite of wear experiments with neat PTFE in various conditions against 304

stainless steel ......................................................................................................................34

2-5 Time lapse images of the wear of unfilled PTFE illustrate the transfer, agglomeration

and growth of PTFE wear debris islands ...........................................................................35

2-6 The wear rates of various PTFE composites with 5 wt. % filler added are plotted

versus their average friction coefficients ...........................................................................36

2-7 A representative plot of the wear rate of a 5 wt. % α-alumina PTFE composite

illustrates the run-in, transition, and steady state behavior of the polymer in sliding .......37

2-8 A summary of Urueña’s transfer film wear study50 ...........................................................37

2-9 Urueña’s transfer film wear data measured using a stainless steel pin50 is compared

to Ye’s similar study57 .......................................................................................................38

2-10 A plot of the total wear rates of a PTFE/α-alumina composite against an array of

surfaces with prescribed angular roughness ......................................................................38

2-11 In a reproduced plot from Harris et al.,76 the total wear rate of the PTFE/α-alumina

composite is plotted against the roughness angle of the countersurface ...........................39

2-12 The wear rates of various PTFE/alumina composites are plotted versus the supplier

designated particle size of the fillers as described by Krick92 and Blanchet63 ...................39

2-13 SEM micrographs and SLS data demonstrate how BET data provided by particle

suppliers may not be an accurate representation of true particle size ................................41

2-14 X-Ray Microtomography adapted from Krick et al.’s 2015 particle size study92 .............42

2-15 TEM micrographs of the running films confirm that nanoscale alumina particles are

present at the sliding surface of the composite ..................................................................43

2-16 XPS spectra and optical images adapted from Krick43,47 indicate chemical changes at

the sliding interface ............................................................................................................44

2-17 The wear rate of a PTFE/α-alumina composite is plotted vs. relative humidity, and

the wear rate of another PTFE/α-alumina composite is plotted vs. pressure .....................44

Page 9: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

9

3-1 Schematic of linear reciprocating tribometer with a flat-on-flat pin configuration, and

a six-axis load cell and several LVDTs for data acquisition. ............................................52

3-2 A schematic of the stripe test run using a PTFE/5 wt. % α-alumina composite pin

against 304 stainless steel with a lapped finish. .................................................................52

3-3 Stripe tests were performed on the etched surfaces of unfilled PTFE and a PTFE/α-

alumina composite .............................................................................................................53

4-1 Wear rates and friction coefficients for each experiment run as a part of the stripe test ...64

4-2 Stylus profiles of transfer films after 1 cycle, 100k cycles, and 1M cycles of

development .......................................................................................................................65

4-3 High resolution XPS of the transfer film in various stages of development .....................65

4-4 Infrared reflectance results from the transfer film on a stainless steel surface ..................66

4-5 A Hamaker solution for the attractive energy between a flat surface and a cylinder ........67

4-6 The carbonyl region of IR spectra of the tribofilms is compared to small molecule

model reactions with perfluorinated carboxylic acids .......................................................68

4-7 The chemical mechanism responsible for PTFE/α-alumina tribofilm formation and

adhesion .............................................................................................................................69

4-8 The wear rate of the PTFE/α-alumina composite and neat PTFE compared to that of

the submerged composite ...................................................................................................70

4-9 The wear rates of the etched and unmodified surfaces of unfilled PTFE and of the

PTFE/α-alumina composite are plotted versus the total sliding distance. .........................70

4-10 Optical images of the transfer and running films of the etched and unmodified

surfaces of unfilled PTFE and of the PTFE/α-alumina composite. ...................................71

5-1 Radical chemistry at the sliding interface proceeds despite mild conditions (low

speed, low nominal contact pressure, and low frictional temperature change) .................74

A-1 A summary of the wear results from PTFE/α-alumina composite experiments against

various metal countersurfaces. ...........................................................................................85

A-2 A summary of the friction results from PTFE/α-alumina composite experiments

against various metal countersurfaces. ..............................................................................86

A-3 A summary of the results of PTFE/α-alumina composite experiments against various

copper containing countersurfaces. ....................................................................................87

A-4 A summary of the results of PTFE/α-alumina composite experiments against lead,

gold and platinum ..............................................................................................................88

Page 10: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

10

A-5 Friction traces from the PTFE/α-alumina composite experiments against three

aluminum alloys and against stainless steel, and optical images of the transfer films ......89

A-6 FT-IR spectra taken within the transfer films formed by the PTFE/α-alumina

composite against three types of steel ................................................................................89

A-7 FT-IR spectra taken within the transfer films formed by the PTFE/α-alumina

composite against three different copper alloys .................................................................90

A-8 XPS spectra taken from the running films formed by sliding the PTFE/α-alumina

composite ...........................................................................................................................90

A-9 FT-IR spectra taken within the transfer films formed by the PTFE/α-alumina

composite against three aluminum alloys ..........................................................................91

A-10 Stylus profilometry traces taken across the center of transfer films formed by the

PTFE/α-alumina composite against three aluminum alloys ..............................................91

A-11 FT-IR spectra illustrate the changes in surface chemistry before and after attempting

to remove the transfer films ...............................................................................................92

A-12 FIB cross sections taken within the wear track and in nascent areas of the three

aluminum alloys tested ......................................................................................................93

A-13 A summary of the wear results of the PTFE/α-alumina composite against a number

of Al 6061 T6 countersurfaces ...........................................................................................94

A-14 Micrographs taken within FIB trenches in the nascent surfaces of five differently

prepared Al 6061 T6 surfaces. ...........................................................................................95

A-15 Hardness and modulus results from nanoindentations performed on the surfaces of

five differently prepared Al 6061 T6 surfaces. ..................................................................95

A-16 A series of images depicting the lapping and polishing process at ALSPI. ......................96

A-17 An EDS spectrum collected from a sample of the solids separated from the lapping

compound taken directly from the lapping wheel used by ALSPI. ...................................97

A-18 Backscattered electron micrographs and EDS maps of the subsurface (~1 mm depth)

of two Al 6061 T6 samples ................................................................................................98

Page 11: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

11

Abstract of Dissertation Presented to the Graduate School

of the University of Florida in Partial Fulfillment of the

Requirements for the Degree of Doctor of Philosophy

TRIBOCHEMICAL INTERACTIONS OF A PTFE/ALPHA ALUMINA COMPOSITE AT

THE SLIDING INTERFACE: A MECHANISM FOR ULTRA LOW WEAR

By

Kathryn Leigh Harris

May 2016

Chair: W. Gregory Sawyer

Major: Materials Science and Engineering

The wear and friction behavior of ultralow wear polytetrafluoroethylene (PTFE)/α-

alumina composites first described by Burris and Sawyer in 2006 has been studied intensively in

the years hence. The mechanisms responsible for the remarkable improvement in wear over

unfilled PTFE are not yet fully understood. The formation of tribofilms on the countersurface

and the running face of the polymer is crucial to the ultra-low wear behavior of the composite on

a metal countersurface. The complete chemical mechanism of transfer film formation and

adhesion, and its role in the exceptional wear performance has yet to be elucidated. Some debate

exists regarding the role of chemical interactions between the PTFE, the filler, and the metal

countersurface. Some have concluded that chemical changes are not an important part of the

ultralow wear mechanism in these materials at all.

A “stripe” test allowed comprehensive spectroscopic studies of PTFE/α-alumina transfer

films in various stages throughout development and led to a proposed mechanism which details

the initiation and adhesion of the tribofilms formed on both surfaces of the wear pair. PTFE

chains (carbon-carbon bonds) are broken mechanically during sliding and undergo a cascade of

reactions to produce carboxylate chain ends that chelate to the metal surface and to the surface of

Page 12: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

12

the porous, friable alumina filler particles. This tribochemical process forms a robust polymer-

on-polymer system that protects the steel countersurface from abrasion, and the polymer surface

from wear. The system is able to withstand hundreds of thousands, and possibly millions of

cycles of sliding with almost no wear of the polymer composite after an initial period of high

wear during run-in.

A mathematical model in support of the hypothesis of mechanical scission of carbon-

carbon bonds in the backbone of PTFE in simple sliding contact is detailed, using the Hamaker

model for van der Waals interactions between polymer fibrils and the countersurface (a cylinder

and a flat surface). The proven necessity of ambient moisture and oxygen is explained in the

mechanism, and model experiments using small molecules further support the assignment of

reactions in the proposed mechanism to the processes at the sliding interface.

Page 13: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

13

CHAPTER 1

INTRODUCTION

The value of tribological advancement is abundantly evident. The estimated cost of wear

and friction processes in the U.S. is estimated to be on the order of 6% of gross domestic product

(GDP)1 or approximately $900 billion in 2011 dollars. A reduction in friction coefficient of a

few percent can have a significant impact on efficiency. In addition to the substantial financial

burden generally attributed to wear and friction processes in power generation and

manufacturing, the lives and comfort of recipients of joint replacement implants depends on the

ability to engineer long-lasting sliding interfaces. An overarching goal of tribological research is

to find materials that will operate reliably at a steady and low friction coefficient while

maintaining an ultra-low wear rate. A comprehensive understanding of tribological materials

will be required as engineers and scientists strive to expand our influence into the frontiers of

space, the human body, and a host of applications demanding performance and durability.

Surface science is an integral element of tribology. Most surfaces must come into contact

with others during their useful lifetime. Surface interfaces are therefore perhaps the most critical

design element in machines and mechanical assemblies, as bulk properties are largely understood

and easily tested. Tribological phenomena at these interfaces are fundamental to system

performance, longevity, and reliability. The two most common metrics for tribological

interactions are friction coefficient and wear rate. The friction coefficient of a system, µ, is

defined as the ratio of the applied normal load to the lateral resistance to sliding. The wear rate,

K, is typically described as a volume loss per unit force per sliding distance, and is commonly

reported in units of mm3/(N∙m). For changes to be made to these metrics, changes must be made

to the surfaces involved. Friction and wear are not material properties, and vary as wildly as do

the applications of the systems in question. Numerous parameters, including choice of materials,

Page 14: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

14

the applied contact pressures, sliding velocity and geometry and environmental conditions must

be considered in tribological design. Tribological interactions are also time-dependent and

frequently may rely on seemingly rare interactions that are a function of mechanics across all

length scales, of chemistry, of physics, and of all manner of materials phenomena.

Commonly, sliding surfaces are protected by the addition of a lubricious separating layer

to postpone damage to or modification of, the working surfaces. Fluid lubricants are useful in

systems operating at high enough speeds to promote hydrodynamic lubrication, but along with

greases are largely less desirable than solid lubricants in applications at higher operating

temperatures and contact pressures, lower speeds, vacuum environments, or that involve

reciprocations. Solid lubricants are in general less sensitive to contaminants, and safer for use in

biological applications. Polytetrafluoroethylene (PTFE) is used in a large number of tribological

applications because of its exceedingly low coefficient of friction.2–6 However, the neat polymer

has a very high wear rate (K~7x10-4 mm3/(N·m)).5–9 PTFE composites have been widely studied

because the inclusion of particulate and polymeric fillers improved the wear rate by several

orders of magnitude.6,8,10–13 Wear abatement of PTFE using fillers of various sizes and chemical

composition appears to be mechanistically unique: PTFE itself has been shown to wear in a

manner quite contrary to other engineering polymers, and the mechanism has been of interest for

decades.5,6,13–15

Page 15: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

15

CHAPTER 2

BACKGROUND

The History of Polytetrafluoroethylene

Polytetrafluoroethylene (PTFE) is a tough thermoplastic which tends to creep and wear in

part due to weak, largely dispersive intermolecular forces. Stronger intermolecular forces are

prevented by the low polarizability of the very strongly ionic C-F bonds present. The relative

weakness of these intermolecular forces contributes to the low surface tension, low wettability

and low friction of the polymer. The inertness and thermal stability of PTFE are also due in no

small part to the strength of the carbon-fluorine bond.16,17

The discovery of PTFE in 1938 by Plunkett was a serendipitous event that occurred while

it was still commonly believed that fully substituted ethylenes could not be polymerized.18

Plunkett noticed that an especially inert, waxy solid had formed under pressure in a valve on a

cylinder of tetrafluoroethylene (TFE), and patented his discovery and polymerization methods in

1941.19,20

Today, PTFE is produced commercially. Pyrolysis of chlorodifluoromethane yields

tetrafluoroethylene, which is polymerized into PTFE using various free radical initiators by

several methods (Figure 2-1). Granular PTFE is produced via aqueous suspension

polymerization and is cut or milled into particles tens of microns in size, while fine powder

resins are manufactured using an aqueous emulsion method and separation by agitation or the

addition of electrolytes. The fine powders (~200 nm) tend to agglomerate as particles (~500 µm),

which may be dispersed in lubricants and paste extruded. PTFE is also available in aqueous

dispersion with non-ionic surfactants that may be used as coatings after application and heat

treatment.21 PTFE radicals are most commonly terminated by combination or by acidic end

Page 16: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

16

groups rather than by disproportionation, or removal of fluorine atoms to form a terminating

carbon-carbon double bond.22

The PTFE molecule itself adopts a helical structure, as well as increase in the C-C bond

angle past the usual 109° (Figure 2-2). This is assumed to be due to the overcrowding of fluorine

atoms that would occur in the zig-zag configuration of hydrocarbons.23 Because of this, the

molecules are cylindrical, or rodlike in shape. Below 19 °C, the helix has a repeat unit that sees a

180° twist per 13 carbon atoms (known sometimes as form II) and packs triclinically, and above

19 °C, the helix expands to rotate 180° per 15 carbon atoms (known sometimes as form IV).24,25

Above 30 °C, a further phase change occurs and the helix becomes more irregular, though it

maintains its lateral hexagonal packing up to its melt temperature, 327 °C.19,23

The high molecular weight of PTFE (ten to hundreds of thousands or even millions of

repeat units)26,27 and its high melt viscosity affect melt crystallization.16 Above the melt

temperature, PTFE forms a gel rather than a liquid, rendering it non-melt processible. As such, it

must be sintered above the melt temperature rather than extruded or molded. Uniquely, PTFE

crystallizes to form banded lamellae with interlamellar regions of non-crystalline polymer, as

proposed by Speerschneider in 1962, rather than the more typical spherulitic configuration of

other polymers28 as illustrated in Figure 2-3. The rate of cooling affects the lamellar thickness,

the percent crystallinity, and thus the density of the polymer. PTFE molecules within the

crystalline lamellae lie perpendicular to the length of a band, and parallel to the striations.26

Despite its exceptional thermal, chemical and frictional properties, neat PTFE is fairly

soft when unreinforced. Though orientation of the molecules and percent crystallinity do affect

the hardness, toughness, and tensile and compressive properties of PTFE somewhat, it is not

Page 17: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

17

capable of supporting high loads without succumbing to creep or failure. Temperature dependent

tensile and compressive strengths of comparable PTFE resins are given in Table 2-1.

The Wear of Neat PTFE

The extremely low friction coefficient of PTFE as well as its chemical inertness, thermal

stability and low vapor pressure contribute to its singular utility as a solid lubricant. Because of

its low resistance to sliding, it has been widely studied by tribologists for decades. However, its

wear rate is markedly higher than that of other neat polymers, and the characteristic ‘island’

shape of the transferred material is fairly unique. The wear rate of neat PTFE is around 10-4

mm3/N∙m, independent of surface roughness or environment.

In 1952, Shooter and Tabor observed that the low friction of PTFE seemed independent

of the nature of the countersurface, and that it did not seem to be a result of lubrication by a self-

mated surface film.2 McLaren and Tabor attributed the speed and temperature sensitivity of

PTFE to viscoelastic origins.29 Two frictional regimes were observed by Makinson and Tabor in

1964.5 They concluded that at low speeds and/or high temperatures, friction was extremely low,

and that thin, aligned films were drawn from the surface asperities during sliding. At high speeds

and low temperatures, friction was increased and transfer of material was greater and thicker.

Furthermore, they were surprised to observe that the adhesion of the transferred material was

quite strong. From they concluded that the low friction coefficient arose from shear failure

between crystallites rather than from lack of adhesion to the surface. Steijn confirmed the

alignment of PTFE fibrils with the sliding direction and concluded that the chain slide past one

another with some ease.30 Pooley and Tabor then attributed the low friction to the lack of side

groups on the PTFE chain, and observed that crystallinity and morphology of the polymer

appeared to have almost no effect on friction and transfer.31

Page 18: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

18

Makinson and Tabor’s conclusion regarding subsurface shear5 and its relation to the

friction and wear regimes of PTFE led to further studies regarding its seemingly contrary wear

mechanism. Tanaka proposed that because no banded structure was visible within even the

thickest wear platelets, gross wear occurred due to destruction of the banded structure proposed

by Speerschneider28 and that fibrillous wear debris was a result of slippage of crystalline slices in

adjacent bands, rather than of the removal of an entire section of a band.8 Uchiyama observed the

motion and growth of transferred PTFE fragments and noted their propensity to combine with

one another and also to transfer back to the sliding surface of the pin.32 In 1992, Blanchet and

Kennedy observed cracks that seemed to be propagating in the direction of sliding, and

concluded that in its high wear regime, PTFE must wear by delamination as subsurface defects

initiated cracks that could not support the shear stress applied during accelerated sliding.6

A suite of wear experiments with neat, compression molded PTFE pins (flat on flat) were

performed in order to illustrate its variable behavior and to observe the apparent mechanisms

first hand (Figure 2-4). All tests were run as 25.4 mm long reciprocations on ~120 nm Ra

stainless steel, and at 50.8 mm/s unless otherwise specified. Two tests against stainless steel in

lab air yielded the expected high speed wear rate of neat PTFE (10-4 mm3/N∙m) and formed the

characteristic patchy transfer film. Submerged in filtered water, the wear rate was nearly an order

of magnitude lower and the adhesion of the wear debris was greatly reduced, though the debris

that floated away from the track varied in size up to visible fractions of millimeters. A self-mated

test resulted in wear rates comparable to the submerged test. At 15 °C (below the 19 °C phase

transition – though no dependence of wear on phase is known) and under impinging dry nitrogen

to prevent condensation of water on the sliding surface, the wear rate of the PTFE pin was as

high as it was during the standard lab air tests, but the wear debris did not form the usual islands.

Page 19: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

19

Instead, the debris was powdery and fine. A test was run at 35 °C (above the 30 °C phase

transition) and 1 mm/s in order to illustrate the speed transition in wear. Indeed, the sample ran

for 50,000 cycles and wore very little. The differences in the wear debris during the higher speed

tests suggest that the islands seen in the wear tracks of unfilled PTFE may require a certain

thermal energy input to form, and that they form from smaller, pre-transferred debris particles at

the sliding interface as sliding progresses rather than always delaminating in large pieces.

Time lapse photos of the transfer film of a neat PTFE pin run on 20 nm Ra nickel at the

same conditions as the previous tests (wear rate was again ~10-4 mm3/N∙m) illustrated the

evolution of the debris islands as the number of sliding passes increased. Images were taken after

every cycle. Every eighth image over the first 112 cycles of sliding are given in Figure 2-5.

Small, loosely adhered PTFE islands formed quickly, and could be seen to move around the wear

track before agglomerating into the large platelets characteristic of the high wear of PTFE. The

wear platelets remained mobile throughout the entire test (1000 cycles), and appeared to grow to

some critical size before being peeled off and removed as debris under the frictional shear stress.

Again it was evident that the large platelets had not delaminated at their final size. Furthermore,

small wear platelets were visible in the optical after just four cycles, so it is unlikely that all

PTFE wear debris is formed by the propagation of subsurface cracks. Indeed, surface plasmon

resonance experiments by Krick et al. confirmed that PTFE can transfer to a countersurface after

a single cycle of sliding.33

Friction and Wear of PTFE Composites

Filler particles and fibers of widely varying size and composition have long been used to

abate the high wear of PTFE, usually at the cost of a slight to moderate increase in the friction

coefficient.6–12 Micrometer-sized metal oxide particles (ZrO2 and TiO2) and glass fibers were

shown to each reduce the wear rate of PTFE by one or two orders of magnitude while

Page 20: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

20

maintaining a relatively low friction coefficient (µ~0.2). It was then supposed that that particles

and fibers of moderate size were more effective at reducing wear than smaller powders and hard

particles.9 At the time, wear reduction via fillers was considered from a largely mechanical

standpoint, and it was presumed that fillers reduced wear by 1) supporting a large portion of the

normal load34,35, 2) arresting the propagation of cracks6,36,37, 3) disrupting or strengthening the

banded structure of the polymer8,9, 4) preventing initial transfer of the polymer9, or 5) modifying

the adhesion of the transfer film.38

Though hard, micrometer sized particles are successful in preventing some wear of

PTFE, they tend to abrade the countersurface, preventing the formation of a smooth transfer film

and creating a roughened sliding interface that exacerbates wear of the polymer. A shift to

smaller particles was therefore desirable to prevent countersurface damage. In recent decades,

various nanoparticles were indeed shown to be capable of reducing the wear rate of PTFE7,39. A

study by Li in 2001 experimented with reducing wear using ZnO nanoparticles40, finding that the

wear was reduced by close to two orders of magnitude, and Chen had similar results using

carbon nanotubes.41 Burris and Sawyer later discovered that the inclusion of 5 wt. % of a

particular α-phase alumina particle reduced wear by an additional factor of 100, a thousand fold

improvement over neat PTFE42, while also maintaining a lower friction coefficient than many

similar composites. It is logical that even low loadings of nanoparticles should have a marked

effect on the properties of the polymer matrix due to their high number density and surface area

compared to particles of greater size. The term “ultralow wear” PTFE composites describes these

exceptional materials with wear rates less than 10-6 mm3/ (N·m).40,42–50 Various other fillers

loaded into PTFE at 5 wt. % were demonstrated by Sawyer’s tribology group to have vastly

Page 21: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

21

differing effects on the wear rate and friction coefficient of the polymer against lapped 304

stainless steel, though none matched the performance of the α-alumina composite (Figure 2-6).

Tribofilms and Ultra-Low Wear

Unfilled PTFE has long been shown to create transfer films of variable thickness and

covered, via mechanisms seemingly contrary to those of other, higher friction polymers.5,30,31 In

1993 Blanchet showed that PTFE transferred to surfaces in sliding, regardless of the chemical

cleanliness of the surfaces or the presence of oxides.10 The low wear of polymer nanocomposites

is in general associated with the formation of thinner, more robust and uniform transfer films on

the countersurface, and running films on the polymer surface.40,42–44,46,51–55 The three regimes of

film formation, run-in, transition, and steady state, are distinguished by morphological changes

to the transfer and running films, and changes in wear rate.45 An example of this behavior is

illustrated in Figure 2-7: the wear rate of a PTFE/alpha alumina composite decreases as sliding

continues and a tribofilms form.

Transfer films for composites of PTFE and α- and Г- phase alumina were observed by

Burris to fill in negative features of the countersurface during the run-in period of higher wear to

provide a self-mated polymer interface that proceeds at a steady state wear rate independent of

the original average roughness of the nascent surface, and the transfer films were seen to increase

in thickness and discontinuity with increasing wear rate of the polymer composites.51 In a set of

experiments that slid transfer films of neat PTFE, Burris also showed that the alignment with

sliding direction of fibrils in the transfer films was conducive to low wear, and that sliding

perpendicular to the aligned direction led to rapid failure of the films.

Because the ultra-low wear behavior of PTFE/alumina composites is always

accompanied by the formation and evolution of visible, reddish-brown tribofilms on both sliding

surfaces, and because the run-in wear of the composites is always higher in the period before

Page 22: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

22

film formation, it follows that the films are crucial to the abatement of wear of the polymer. A

comprehensive study of some of the mechanical properties of the tribofilms was conducted by

Krick et al. in 2014, who documented the color change, wear transitions, the hardness and

modulus of the running films in various stages of formation, and the depth of the effects of the

running film on hardness and modulus.47 They confirmed that the transition from high to low

wear occurred in conjunction with the formation of the tribofilms, that the hardness and modulus

of the running film increased as wear progressed, and that these increases were greatest at the

near surface, but were detectable at indentation depths up to around 250 nm.

Following Krick et al.’s analysis of the running films, Urueña et al. studied the wear

resistance of the PTFE/α-alumina transfer films themselves by performing microtribometer

experiments on transfer films in various stages of development.50 A single PTFE/ α-alumina pin

was used to run a “stripe test”56 that decreased stroke length after every mass measurement in

order to expose areas of transfer film in seven intervals from one thousand to one million cycles.

Microtribometer tests using a stainless steel ball of radius 1.5 mm measured friction

coefficients in each of the exposed areas. The transfer film patches were said to be worn through

after a number of microtribometer cycles corresponding to a crossover in friction coefficient

from low and consistent to high and erratic. Stylus profilometry was used to measure the initial

thickness of each of the exposed patches, and the wear volume used to calculate the wear rate of

the transfer film in each test was determined by Scanning Electron Microscopy (SEM). The

results of these tests indicated that the transfer films did become more robust as development

progressed, but that the films themselves were not necessarily a low wear material. The wear

rates of the films were several orders of magnitude higher than those of the composite pins. The

results of the study are summarized in Figure 2-8.

Page 23: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

23

Soon after these results were published, a study by Ye expounded upon them by

measuring the adhesive and cohesive strength of PTFE/α-alumina transfer films in tension, as

well as the sensitivity of the wear rate of the films to the surface energy of the probes used in

testing.57 Using the tensile strip methods of Agrawal and Raj and the Tresca failure criterion,58–60

Ye showed that the ratio of adhesion to shear strength of the films increased with increasing

transfer film development, and added that a film could not be considered to be adhesive until the

ratio was greater than one – a condition which was quickly reached by the films after several tens

of meters of sliding. Furthermore, it was discovered that against surface probes of low surface

energy (PTFE, HDPE, other polymers) that the wear rates of the films were in fact

extraordinarily low (10-8 to 10-9 mm3/N∙m), comparable to the steady state wear rates of the

composites once the self-mated films condition is established. Ye’s results are compared to

Urueña’s in Figure 2-9.

The omnipresence of tribofilms on the sliding surfaces of ultra-low wearing PTFE/α-

alumina nanocomposites has led to a number of forays into the mechanical changes undergone. It

is evident that the films play a crucial role in maintaining the low wear behavior of the system.

Nanoparticle fillers especially seem to facilitate the long life of the films as they are too small to

cause abrasive damage or third body wear within the contact.

The Effects of Surface Roughness

While counterface asperity shape may in some cases govern the friction coefficient,

asperity size and radius of curvature may govern wear.36,61,62 Blanchet et al. showed that surface

roughness and filler concentration strongly affect the wear rate of PTFE/alumina composites, and

that it transitioned from K~10-7 mm3/(N·m) to K~10-4 mm3/(N·m) past a critical roughness

threshold for a given composite composition.63 For the PTFE/alumina composite considered in

the following experiments, the roughness threshold was reported to be ~ 5 µm. Franklin et al.

Page 24: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

24

showed that surface roughness aligned perpendicular to the sliding direction increased the wear

of a POM-20% PTFE composite with increasing surface roughness,64 and Friedrich showed that

unfilled PEEK sample wore more with increasing roughness aligned parallel to the sliding

direction.65

Recently, studies of the ultralow wearing PTFE/α-alumina composite system have

investigated the characteristics of the tribofilms formed during sliding rather than of the polymer

composites themselves. Third bodies (debris, transfer films and running films) affect the friction

and wear of solids, as has been documented for many material systems.53,66–75 The following

study by Harris et al. in 2015 investigated the link between the ability of third bodies to

accumulate on a countersample and the transition to ultra-low wear sliding.76

Stainless steel countersamples were prepared with aligned grooves defined by the angle

of the grooves relative to sliding direction, referred to as the “roughness angle”, θ. The roughness

angle varied from 0° (parallel to the sliding direction) to 90° (perpendicular to the sliding

direction). Directional roughness was prescribed using 600 grit silicon carbide sandpaper

adhered to a thick sheet of polydimethylsiloxane (PDMS), which was mounted to the stage of a

linear reciprocating tribometer.77 The stationary countersample was mounted to a jig that allowed

adjustment relative to the sliding direction in 5° increments. The aligned countersample was run

at 250 N for 200 reciprocating cycles, resulting in an average roughness (Ra) of 140 nm with a

standard deviation of 28 nm across all countersamples and an average groove width of ~60 µm.

No sample exceeded 220 nm Ra. A lapped finish (Ra ~150 nm) and a highly polished mirror

finish (Ra ~ 20 nm) were used as controls. All countersamples were washed with soap and water,

sonicated in methanol, and allowed to dry in laboratory air prior to experiments. Analysis of the

Page 25: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

25

surface features of the metal countersamples was performed using a scanning white light

interferometer. Scan data were averaged over ten locations on each countersurface.

The wear rate of the PTFE/α-alumina composite was evaluated in sliding against six

different roughness angle surfaces, the lapped surface, and the polished surface. The total wear

rates at each mass measurement for each of the countersurface treatments is plotted in Figure 2-

10. All of the samples experienced transient wear processes, including a transition from an

initially higher wear rate to a lower final wear rate (run-in). The run-in transition in wear rate

was more dramatic for materials that reached ultralow wear rates (θ = 90°) than for those that

had high total wear rates (θ = 0°). Generally, the total wear rate and the steady state (final) wear

rate decreases as the roughness angle (θ) increases, producing the highest wear rate when sliding

parallel the roughness features and lowest wear when sliding perpendicular to the ridges. At

roughness angle 0°, the PTFE/α-alumina composite had a high steady state wear rate (9.6 × 10-6

mm3/(N·m)), and never reached ultra-low wear over the duration of the test. The steady state

wear rate for PTFE/α-alumina on the 90° countersample, however, was extremely low (K ~ 3.1 ×

10-8 mm3/(N·m)), lower than the lapped and the polished controls. The wear rate in the parallel

direction was around 300 times greater than the wear rate in the perpendicular direction.

Optical images and scanning electron micrographs of the wear tracks revealed transfer

film morphology and coverage. The 0° countersample developed a dark transfer film at the ends

of its stroke, but the center appeared largely bare. Electron micrographs of the zero degree

transfer film showed flaky patches of wear debris or transfer that appeared to be resting on top of

the roughness ridges. The 90° test formed a fairly uniform transfer film that covered most of the

sliding surface. The 45° film had characteristics of both the 90° and 0° countersamples and did

not develops darker orange regions or patches.

Page 26: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

26

PTFE/alumina transfer films adhere strongly to steel countersurfaces. In the context of

debris mobility, the longer a third body (debris or tribofilm) remains in sliding contact, the

higher the probability of transfer film adhesion. By designing surfaces to retain or eject debris,

we can control the amount of mechanical work done on debris particles for a given tribological

system. On countersamples with roughness features aligned in the sliding direction, debris is

easily shed and wear rates remain high. When roughness features are perpendicular to the sliding

direction, debris may be trapped. This allows for increased sliding cycles over a given third

body, resulting in increased opportunities for adhesion and ultimately ultralow wear rates. On the

90° surface, retention of debris and formation of a stable transfer film appeared to produce steady

state wear rates significantly lower than on randomly rough surfaces of comparable Ra. Of

course, while the 90° surface had the largest capacity to retain debris, it required a greater run-in

wear volume to fill in the negative surface features before ultra-low wear was reached. The

results of this study suggested that fillers which abrade and damage to the metal surface create a

surface that promotes third body rejection and thus does not facilitate ultra-low wear.

Debris mobility may be modeled in several ways (Figure 2-11). Godet explored factors

that could affect the recirculation or ejection of third bodies from contact.74 Contact shape,

roughness and third body rheology effects were proposed based on this work78–80. Thereafter,

optical and chemical in situ studies by Wahl, Singer, and Chromik looked into the role of third

bodies.81–84 The study described in this chapter was specifically designed to control the mobility

of debris and its ejection from the wear track, and provided a simple model for debris mobility

based on the number of obstacles in the path of the debris. For the surfaces with aligned

roughness, these obstacles were the ridges. In a given sliding cycle, a mobile debris particle

crosses a number of ridges proportional to sin . A simple model considered the wear resistance

Page 27: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

27

to increase with the number of ridges crossed during the stroke – the wear rate then decreased

with theta as1 sin .

A slightly more complex approach considered the local forces on debris and third bodies.

The local friction force is composed of a reaction force against the ridges and a debris migration

force along each ridge. With increasing θ, the migration force along the ridge decreases and the

reaction force increases. The ratio of the migration force to the reaction force resembles a local

friction coefficient. Wear rate appears to scale with the mobility factor, meaning that higher

mobility of debris leads to a higher wear rate. Real behavior of these wear systems depends on

many factors aside from debris mobility. Debris and transfer film retention likely depends on a

critical shear strength of the interface between countersurface and third bodies, dependent on the

size of surface features and roughness angles. At the critical shear stress, material could migrate

out of the grooves, increasing wear, resulting in sparse transfer films as seen on the higher

wearing samples.

The Effects of α-Alumina Particle Morphology

Granular PTFE 7C with 1-5 vol. % alumina particles added have been shown to reduce

the wear of the polymer to 1x10-7 mm3/(Nm) or lower [5-24].6,7,9,11,35,36,42,43,85–91 This was first

described by experiments using three different alumina fillers of manufacturer designated

particle sizes of “44 nm”, “80 nm” and “500 nm” [17, 19-24].35,42,43,51,63,89,90 The “80 nm”

alumina achieved steady state wear rates near 8x10-7 mm3/(N∙m) - lower than the other

composites by a factor of 100. No correlation between wear and particle size was observed.

Later, it was determined that the “44 nm” particles were spherical delta-gamma phase alumina,

and the others were alpha-phase.51 McElwain et al. and Blanchet et al. reported that the wear rate

of PTFE with 2.9 vol. % (~ 5 wt. %) alpha-phase alumina nanoparticles was two orders of

Page 28: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

28

magnitude lower than with “microparticles”,35,63 and were also relying on the supplier designated

particle size.

One hypothesis regarding the effect of the addition of inorganic particles is that they

somehow interrupt the subsurface crack propagation of PTFE.35 Another hypothesis states that

wear reducing additives alter the crystalline structure of the PTFE into something tougher,

although so far there is no definitive evidence for such a mechanism.6,10 It was also proposed that

the banded structure of PTFE was disrupted, accounting changes in wear performance.8,9 In the

following summarized study,92 five alumina powders were used as fillers at 5 wt. % in Teflon®

PTFE 7C:

Alumina A-Alfa Aesar α-phase alumina powder with a supplier-specified approximate

particle size of 0.35 to 0.49 µm (Stock #42573, 99.95%)

Alumina B - Almatis calcined α -phase alumina powder (grade A 16 SG, 99.8%) with a

supplier-specified typical d50 particle size of 0.5 µm

Alumina C - Alfa Aesar α -phase alumina powder (Stock #44652, 99%) with a supplier-

specified approximate particle size of 60 nm, and confirmed to lead to ultra-low wear42

Alumina D - Alfa Aesar α -phase alumina powder (Stock #44653, 99%) with a supplier-

specified approximate particle size of 27-43 nm, also previously reported to produce

ultralow wear PTFE composites35,42,43,46–51,63,89,90,93–96

Alumina E - Nanostructured and Amorphous Materials, Inc. alpha-phase alumina powder

(Stock # 1015WW, 99.5%, according to the manufacturer, mostly alpha-phase with 5-

10% gamma phase) with a supplier-specified approximate particle size of 27-43 nm,

another filler confirmed to result in ultra-low wear of PTFE35,42,63,91

The wear rates achieved by PTFE samples filled with each of these types of alumina, as

well as previously observed values thereof are given in Figure 2-12. The results for PTFE filled

with Alumina A, C, and D agreed with Blanchet’s study. The composite filled with Alumina B

behaved unexpectedly given its supplier designated particle size was equal to that of Alumina A.

The publications mentioned before observed a large transition in wear rate appearing

above about 100 nm in particle size as reported by the vendor. Particle size distribution

Page 29: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

29

measurements were conducted by Static Light Scattering (SLS), by the Brunauer, Emmett and

Teller method (BET), Scanning Electron Microscopy (SEM). The particle size measurements

supplied did not match the particle sizes measured independently. Results are detailed in Table

2-2 and are listed in increasing order by median particle size as determined by SLS. It was

assumed that the particle sizes provided by the vendors were measured using the BET method,

which assumes that the particles are spherical, dense, and monodisperse. Many of the alpha-

phase alumina particles used in these studies were discovered to be irregular in shape, and porous

by SLS and SEM, as demonstrated in Figure 2-13. For this reason, the assumptions for BET

particle size calculation would have been invalid.

In conjunction, SLS, BET surface area measurement, and microscopic examination of the

particles suggested that surface area and porosity of the particles increases from Alumina A to E.

It is commonly assumed that the large surface area of nanoparticles is one of the main sources of

wear reduction. However, in this case the internal porosity is not accessible by the PTFE

polymer chains in the bulk of the part. Furthermore, the transfer film topography results seemed

to contradict the SLS particle size measurements: The smallest particles by SLS, Alumina A,

resulted in abrasion of the steel countersample and wear scars significantly deeper than the

measured particle size. Alumina E had the largest particle size by static light scattering, but its

composite had the lowest wear rate and formed films that were significantly thinner than the

alumina particle size. Even so, a recent XPS study reported an alumina concentration between 1

and 5 at. % at the transfer film surface.48 It was thus hypothesized by the authors that the large

particles of Alumina E were friable and prone to breaking up due to mechanical stresses at the

sliding interface, preventing damage to the metal and the transfer film. Indeed, small, hard,

Page 30: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

30

inorganic fillers in low wearing PTFE have been shown to accumulate at the polymer wear

surface and increase the abrasive wear of the steel surface.97

X-ray microtomography was used to investigate the near surface region of the

PTFE/Alumina E polymer composite after wear testing. Images acquired showed evidence that

alumina particles smaller than the particle size measured by SLS accumulated at the wear

interface, and also revealed that micrometer-sized particles remained in the bulk, as shown in

Figure 2-14. This result was strongly in support of the hypothesis that large friable particles in

the bulk broke up during sliding and that the resulting smaller particles built up at the interface.

The X-Ray Microtomography results were further confirmed by TEM micrographs,

shown in Figure 2-15. The TEM images are consistent with the X-ray microtomography results

in that both techniques give evidence of occasional well-distributed microscale filler particle, and

in that a distinctly different region within the top few micrometers of the worn polymer surface

contains fine scale features that Energy Dispersive Spectroscopy (EDS) in both the TEM and

SEM indicated are alumina.

This study posited that the best alumina particles for ultralow wear PTFE composites are

porous and micrometers in size, rather than dense nanoscale spheres, and that supplier-

designated particle sizes, likely based on BET surface area measurements, were unreliable. The

micrometer sized mesoporous-like filler materials are comparable to nanoparticles in surface

area, and have nanometer-scaled features. In general, wear rate of the composite decreased with

increasing BET surface area to average particle size ratios. It was agreed that multi-scale fillers

are ideal for such large wear reductions. The porosity of the alumina particles rendered them

mechanically friable, the micrometer-scale of the particles likely reduced subsurface

Page 31: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

31

delamination of PTFE composites in the bulk to near surface region,47–49 and the nanometer-scale

alumina particle fragments were still assumed to stabilize and reinforce tribofilms.

Tribochemistry

Several studies have correlated chemical degradation of PTFE with changes in

mechanical properties98–101 and color.102 Chain scission and defluorination due to degradation

were suspected to lead to the presence of CF3, carbon-carbon double bonds, branched carbon

structures, COF and carboxylic acids groups, and crosslinks.103–107 XPS analysis by Scott Perry’s

group at the University of Florida found α-alumina concentrations in transfer films comparable

to the composites, and higher oxygen content than predicted which suggested oxidation of PTFE

occurred during sliding. Furthermore, a correlation of friction coefficient with oxygen content of

the transfer film was observed, though it could not be said which caused the other, or that

causation was even present. Based on this correlation, it was predicted that oxygen content was

lower at the ends of the wear tracks, where friction is lowest.108 Burris then suggested that

enough energy is absorbed in the thin layers during sliding to initiate low probability chemical

events. Further XPS analysis by Professor Perry’s group at the University of Florida indicated a

chemical change within the tribofilms as evidenced by the formation of a new peak at (288 eV)

and was expounded upon by Krick (Figure 2-16).43

Over several studies, Blanchet showed that e-beam irradiated PTFE composites

demonstrated increased wear resistance.109,110 Burris also showed that a chemically etched

unfilled PTFE sample was 100 times more wear resistant and had 10% higher friction than the

untreated polymer before the etched surface wore through. Furthermore, the etched polymer

displayed a peak at 288 eV corresponding to the peak seen by earlier XPS studies of transfer

films.108 This compelling evidence for tribochemical contribution to the ultra-low wear of a 5 wt.

% PTFE/α-alumina led Krick et al. to conduct a series of wear experiments in varied

Page 32: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

32

environment to elucidate the effect of oxygen and water on the system.43 The composites wore

more with decreasing relative humidity, and performed slightly better in environments

containing oxygen than in those containing pure nitrogen. However, some oxygen was present

during all test below their detectable limit of 100 ppm. The wear of the composite was highest

submerged in water. They concluded that water is necessary for the ultra-low wear behavior of

PTFE/α-alumina composites, and that oxygen is likely necessary. The reduction in wear rate in

the presence of water suggested that a secondary wear mechanism was reduced by the formation

of thin, protective tribofilms. Tribofilms associated with ultra-low wear rates did not form in the

absence of water. Soon thereafter, Pitenis et al. conducted a similar suite of tests on the same

composite, varying the level of vacuum of the test environment.46 The wear rate of the composite

was highest in the lowest pressure system (~10-6 Torr), where a visible transfer film did not form,

and highest at 5 Torr, where it was suggested that the water to oxygen ratio was most favorable

for film formation. At the lowest pressure tested, oxygen was likely mostly absent, leaving a

mostly water environment, which was insufficient to promote film formation or ultra-low wear.

Figure 2-17 compiles Krick’s environmental data and Pitenis’ vacuum data, demonstrating the

dependence of wear behavior of the PTFE/α-alumina composite.

The importance of ambient species, in particular oxygen and water, to the longevity of

the composite suggested that the chemical mechanism for tribofilm formation must likewise

depend on their presence. The identification of this mechanism was of interest in order to fully

comprehend the role of the tribofilms in the ultra-low wear system created by the PTFE/α-

alumina composite, and would introduce the possibility of advancements in composite

production that could lead to increased part life by further facilitating the reactions involved.

Page 33: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

33

Figure 2-1. The polymerization reaction of PTFE. Chlorodifluoromethane yields

tetrafluoroethylene and HCl after pyrolysis. Free radical polymerization of

tetrafluoroethylene creates the PTFE molecule.

Figure 2-2. The PTFE molecule is helical in structure. The twist of the helix is dependent on

temperature (two of its phase transitions occur at 19 and 30 °C).

Page 34: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

34

Figure 2-3. PTFE crystallizes to form lamellae rather than spherulites. An illustration of the

banded structure of PTFE shows large lamellae running left to right with smaller

lamellae lying perpendicular between them. PTFE molecules within the crystalline

lamellae lie parallel to the striations.26

Figure 2-4. A suite of wear experiments with neat PTFE in various conditions against 304

stainless steel illustrates the speed dependent wear transition and the wear behavior of

the self-mated and submerged polymer.

Table 2-1. Tensile and compressive yield strengths of unfilled PTFE as reported by *Rae111 and

**DuPont112.

Page 35: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

35

Figure 2-5. Time lapse images of the wear of unfilled PTFE illustrate the transfer, agglomeration

and growth of PTFE wear debris islands. Nascent steel is shown at the top left, cycle

1 in the second image, and cycle 8 in the third. Images proceed in 8 cycle intervals

thereafter, concluding with cycle 112 (bottom right). A scratch on the steel surface is

visible in the upper left corner of each image, and proves that each image was taken

in the same location as the wear debris evolved. Photos taken by author in Spring

2014.

Page 36: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

36

Figure 2-6. The wear rates of various PTFE composites with 5 wt. % filler added are plotted

versus their average friction coefficients. The engineering ideal lies at the lower limit

of wear and friction coefficient.

Page 37: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

37

Figure 2-7. A representative plot of the wear rate of a 5 wt. % α-alumina PTFE composite

illustrates the run-in, transition, and steady state behavior of the polymer in sliding.

Once very low wear rates are reached, they are maintained for hundreds of thousands

of sliding cycles.

Figure 2-8. A summary of Urueña’s transfer film wear study results that suggested the transfer

films formed by PTFE/α-alumina composites are not intrinsically low wear materials,

but are instead an integral part of a self-mated low wear system. This plot is a

reproduction from Urueña’s 2015 study.50

Page 38: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

38

Figure 2-9. Urueña’s transfer film wear data measured using a stainless steel pin50 is compared to

Ye’s similar study which used a lower surface energy High Density Polyethylene

(HDPE) probe to wear away the transfer films.57 Wear rates of the composite pins

used to create the transfer films and the wear rates of the transfer films themselves are

superimposed. Ye’s study found that the PTFE/α-alumina transfer films were

exceptionally resistant to wear against low surface energy polymer probes.

Figure 2-10. A plot of the total wear rates of a PTFE/α-alumina composite against an array of

surfaces with prescribed angular roughness is reproduced from Harris’ et al.76 The

total wear rates decreased as the surface roughness angle approached 90°

(perpendicular to the sliding direction).

Page 39: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

39

Figure 2-11. In a reproduced plot from Harris et al.,76 the total wear rate of the PTFE/α-alumina

composite is plotted against the roughness angle of the countersurface used in each

test. The two models proposed are superimposed for comparison.

Figure 2-12. The wear rates of various PTFE/alumina composites are plotted versus the supplier

designated particle size of the fillers as described by Krick92 and Blanchet.63 Krick’s

study goes on to show that the particle sizes assigned by the suppliers were largely

inaccurate.

Page 40: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

40

Table 2-2. A summary of the particle size results, which vary by method, and the wear rate of the

PTFE composite containing them. The supplier designated particle sizes, BET

particle sizes (assuming dense spherical particles), and SLS particle sizes of each of

the types of alumina studied, and the wear rates and friction coefficients of the

composites molded at 5 wt. % of each of the types of alumina. A large discrepancy

was observed between the particle sizes estimated by the BET method and by SLS.

The lowest wear rate was achieved by the composite filled with alumina E, which had

the smallest particle size according to the supplier and according to BET, but by far

the largest particle size as measured using SLS. This table is comprised of data from

Krick’s 2015 particle size study.92

Page 41: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

41

Figure 2-13. SEM micrographs and SLS data demonstrate how BET data provided by particle

suppliers may not be an accurate representation of true particle size. (a) SEM

Micrographs of alumina A showed that the particles were close in size to the

supplier’s designation. (b) SEM micrographs of alumina E illustrate the comparative

dimension of the particles to the much smaller size assigned by the supplier. (c) SLS

data for each of the alumina types confirms that the particle size distributions are in

some cases very different than the designations. Images and data are adapted from

Krick’s 2015 particle size study.92

Page 42: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

42

Figure 2-14. X-Ray Microtomography adapted from Krick et al.’s 2015 particle size study92

shows the very small alumina particles that built up in the running film during sliding,

as compared to the more dispersed particles in the bulk.

Page 43: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

43

Figure 2-15. TEM micrographs of the running films confirm that nanoscale alumina particles are

present at the sliding surface of the composite. The large, friable alumina E particles

break up at the interface and support the running surface of the polymer. Micrographs

courtesy of DuPont.

Page 44: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

44

Figure 2-16. XPS spectra and optical images adapted from Krick43,47 indicate chemical changes

at the sliding interface. (a) XPS analysis of the bulk polymer and the tribofilms

formed during sliding indicate the creation of new chemical species during sliding.

(b) Evidence of chemical alteration of the surface is seen in the color change

undergone by the running surface of the composite.

Figure 2-17. The wear rate of a PTFE/α-alumina composite is plotted vs. relative humidity, and

the wear rate of another PTFE/α-alumina composite is plotted vs. pressure. Humidity

studies used alumina C as described above and are adapted from Krick.43 Vacuum

studies used alumina E as described above, and data is adapted from Pitenis.46

Page 45: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

45

CHAPTER 3

METHODS AND EXPERIMENTATION

Materials and Sample Preparation

A PTFE/α-alumina composite was made using DuPont Teflon® PTFE 7C resin as the

matrix. The polymer was filled with 5 wt. % α-phase alumina (Nanostructured & Amorphous

Materials Inc., Stock#: 1015WW,45,46,48,96 (alumina E from the particle morphology study

described previously). The dry PTFE powder was combined with the alumina filler and

submerged in extra dry isopropanol to a total volume of 200 mL. The mixture was sonicated

using an ultrasonic horn (Branson Digital Sonicator 450 with a titanium tip at 40% amplitude for

three, one minute long sessions with 45 seconds rest between each). The dispersion was allowed

to dry completely in a fume hood for seven days. The dried powder mixture was then

compressed in a hydraulic press to approximately 100 MPa in a 440C stainless steel cylindrical

mold before being sintered in an oven. The heating process ramped upwards at 2 °C/min to 380

°C, where it was held for four hours and then cooled to room temperature. The sample was then

machined into a pin (6.3 x 6.3 x 12.7 mm). The square running faces of the pin were polished

with 800 grit silicon carbide sandpaper to an approximate average roughness (Ra) of 100 nm,

measured using a scanning white light interferometer. Finally, the sample was sonicated for 30

minutes in methanol and allowed to dry completely before testing.

The countersample used was a flat, rectangular (115 x 25 x 3.7 mm) plate of 304 stainless

steel finished by lapping (Ra ~ 150 nm) by Alabama Specialty Products Inc., the standard

running surface used in previous experiments with the PTFE/alumina composites.43,46,48,51,89,113

The countersample was cleaned with soap and water, rinsed with methanol, and allowed to dry

prior to experiments.

Page 46: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

46

Tribometer and Wear Test Design

Wear testing was performed using a linear reciprocating tribometer (Figure 3-1) with flat-

on-flat sample geometry as described by Schmitz.77 A six-axis force transducer measured the

normal and frictional forces continuously throughout each test. The running face of the polymer

sample was loaded against the stainless steel countersample at 250 N, or a nominal contact

pressure of around 6.3 MPa (around 50% of the compressive yield strength of neat PTFE). The

contact pressure can be approximated using this nominal calculation because the polymer sample

wears into conformity with the flat countersurface within the first thousand cycles. Pre-existing

asperities are under much higher local contact pressures and wear away quickly. The steel

countersample was mounted to a linear ball-screw stage which reciprocated at 50.8 mm/s.

A “stripe test” 48,50,56 was performed in which the reciprocating stroke length decreased as

the number of sliding cycles increased at predetermined cycle intervals (Figure 3-2). The initial

reciprocating stroke length was 88.9 mm and was reduced by 10.2 mm after each experiment.

The final test was 27.9 mm long. This pattern of transfer film formation isolated and preserved

areas of the polymeric film over the various stages of development. These exposed areas allowed

physical and chemical analyses of the steps involved in transfer film evolution.

Wear Rate Calculation, Friction Measurements and Uncertainty

The wear rate, K, of each polymer sample is described as the volume lost per unit force

and distance and is recorded in units of mm3/N∙m (Equation 3-1), as described by Archard114.

The density, ρ, of each pin is calculated prior to testing using its measured dimensions, L, W1 and

W2, measured using clean calipers, and its initial mass, mi, measured using a Mettler Toledo scale

(Equation 3-2). Volume loss, ΔV, is then the mass loss, Δm, divided by the density (Equation 3-

3). Wear rates were calculated both as total volume lost per total distance slid (total wear rate),

and as volume lost per test per distance slid per test (test wear rate). The total wear rate includes

Page 47: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

47

the run-in period of relatively high wear before the onset of ultra-low wear after stabilization of

the tribofilms and is calculated by considering the volume lost and distance traveled over all

previous tests. The test wear rate differentiates between early, higher wear rates, and the ultra-

low minimums reached after many sliding cycles by considering only the volume lost and

distance traveled during a single test. The steady state wear rate is defined as the wear rate

reached and maintained after the run-in period.

n

V

F D

(3-1)

1

i

i i

mLWW

(3-2)

mV

(3-3)

Volume calculations using vertical displacement measurements are generally less

accurate than those based on mass loss due to the possibility of creep of the polymer, or the

formation of transfer films of non-negligible thickness. The normal load is monitored throughout

the duration of each experiment. The mean value of the force, Fn, is then used to calculate the

wear rate of the polymer over the period for which it is averaged. The lateral displacement, d, of

the reciprocating stage was similarly recorded throughout each test and used to calculate the total

distance traveled, D.

The error in each measured quantity is as follows: u(m) = 0.02 mg, u(L) = 0.02 mm,

u(Fn) = 0.2 N, and u(d) = 0.04 mm/mm. The error in the total distance traveled is

( ) ( )u D D u d , and the error in the mass loss is ( ) 2( ( ))u m u m , according to the law of

propagation of uncertainty. Further propagation of these errors leads to expressions for the

uncertainty in the measured volume loss, V, as given by (Equation 3-4), as well as for the product

Page 48: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

48

of the average normal load and total distance traveled (the denominator of the expression for the

wear rate, K), given by Equation 3-5.

2

2 2

2 2i i

2

i i

V ΔmVu(ΔV) = u(Δm) + u(m)

m m

2

2 2 2 2

i 1 i 2 1 2

i

Δm+ u(L) (LW ) +(LW ) +(W W )

m

(3-4)

2 2 22 2

n n nu(F D) = u(F ) (D) + u(D) (F ) (3-5)

The uncertainty in ΔV and in Fn∙D is used to perform 1000 Monte Carlo simulations of

each wear test. In each simulation, possible values for each parameter are generated randomly

within plus or minus three times the value of the calculated uncertainty. The wear rate is then

calculated for each of the simulations, and an average and a standard deviation thereof is

obtained for each experiment.

Friction coefficients were calculated as the quotient of the average frictional force (measured

continuously throughout each test, and averaged at the end of each test, and the average normal

force77. The associated uncertainties in the friction coefficient and wear rate calculations are

described in greater detail in Schmitz et al.77,115

Stylus Profilometry

Stylus profilometry was used to map transfer film topography (KLA Tencor P-16 with a

2 µm radius probe and 5 mg load). Fifty 12 mm line scans were acquired to map 2 mm x 12 mm

sections in each exposed area of transfer film development. The nearby bare stainless steel

surface was used as a zero reference to shift the profiles according to a common baseline. The

average thickness of each exposed section of the film was calculated as the difference between

Page 49: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

49

the nascent metal zero reference and the center of each film section scanned in the 2 x 12 mm

sections described above.

Small Molecules Experiments

Pressed polymer disks were created at DuPont for IR comparison to the tribofilms formed

during the wear test. The same α-alumina used to create the PTFE composite pin was pre-dried

and heated in a flask in an oil bath at 150 °C and 20 milliTorr for five hours. The flask was

backfilled with dry nitrogen and transferred into a nitrogen-atmosphere drybox. In a separate

flask in the drybox, 10 grams of the alumina were combined with tridecafluoroheptanoic acid

(TCI America, 98.0%, 0.23 g), and dry isopropyl alcohol (Fisher Scientific, 30 mL). The mixture

was stirred in the drybox, transferred to a fume hood and reduced in vacuo (150 Torr) in a water

bath at 28 °C in order to minimize vaporization of the acid. The acid/alumina mixture was added

at 5 wt. % to 10.0 grams of pre-dried PTFE and mixed for 18 hours on a roller mill. The polymer

mixture was cold-pressed under 2.5-3 tons pressure at ambient temperature into 13 mm circular

disks using approximately 100 mg of material in a hydraulic press. The resulting PTFE films

were around 355 µm thick.

X-Ray Photoelectron Spectroscopy

X-ray photoelectron spectroscopy (XPS) was used to analyze each area of exposed

transfer film. XPS was performed at DuPont using a Physical Electronics Quantera Scanning

ESCA Microprobe with a focused (100 um) monochromatic Al K-alpha X-ray (1486.6 eV) beam

at 18 kV and 100 W. The electron energy analyzer was operated in constant energy mode with a

pass energy of 55 and a 0.2 eV step size between points for high resolution spectra (energy

resolution of the system was approximately 0.84 eV using the Ag3d5 peak). Atomic percent

concentrations were normalized to 100%. A dual electron and argon ion beam system was used

for charge compensation.

Page 50: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

50

Infrared Spectroscopy

The infrared spectra used to analyze the metal surfaces and tribofilms were obtained by

collaborators at DuPont using a Nicolet 6700 FT-IR spectrometer with a Continuum Microscope

(Thermo Fisher Scientific) in reflectance mode. The area of analysis was a 100 µm square.

Background spectra were collected away from the fluoropolymer wear track in a clean area of

the steel countersample. Transfer film spectra were obtained by reflectance at three spots along

the centerline at the midpoint of each of the seven exposed transfer film areas. This triplicate

analysis yielded consistent results for each of the areas inspected.

The cumulative wear debris at the end of the one million cycle region was analyzed using

attenuated total reflectance infrared (ATR-IR) on a Golden Gate (Specac) horizontal diamond

ATR unit. Background spectra were collected with a clean diamond surface. Transfer film

residue spectra were collected after the film areas were analyzed and removed from the diamond

surface, but before the crystal was cleaned with ethanol. No pressure was applied from the

clamping device for residue spectra collection. Spectra were corrected for the wavelength

dependence of penetration depth to closely resemble transmission spectra.

Transmission spectra of the pressed polymer films (PTFE 7C/α-alumina/C6F13COOH and

PTFE 7C /α-alumina) were obtained at DuPont using a Nicolet Magna 560 FT-IR spectrometer

(Thermo Fisher Scientific). A background spectrum was collected using an empty film card of

the same type used to mount the films. Acquired spectra were converted to absorbance spectra

for comparison. The C-F overtone peak near 2365 cm-1 was used as a guide to detect changes in

this region, after spectral subtraction of a PTFE control film. Due to the intense C-F stretch

region of these thick disk samples, the region below ~1320 cm-1 was distorted.

Page 51: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

51

Etched PTFE Tests

Unfilled DuPont Teflon® PTFE 7C and a PTFE 7C/α-alumina composite filled with 5

wt. % α-phase alumina (Nanostructured & Amorphous Materials Inc., Stock#: 1015WW,45,46,48,96

(alumina E from the particle morphology study described previously) samples were prepared as

described above. These samples were machined into pins approximately 6.3 x 6.3 x 12 mm in

dimension and polished using 800 grit SiC sandpaper. One each of the filled and unfilled

polymer samples were then etched using Fluoroetch® (Acton Co.), which modifies and oxidizes

the polymer surface so that it contains carboxylic acids116, and caused the exposed surfaces to

turn a dark, chocolate brown. The samples were not polished, in order to preserve the surface

treatment on the running faces of the pins. Before testing, the pins were sonicated in methanol

for 30 minutes and allowed to dry overnight. A stripe test48,50,56 was performed using the etched

surfaces of the PTFE and the α-alumina composite (Figure 3-3). The running films of each of

these samples were removed, and a stripe test was then performed using each of the inner

surfaces of the unfilled PTFE and the α-alumina composite. The PTFE/α-alumina composite ran

for 1k, 10k, and 100k cycles, and the unfilled PTFE samples ran for 10, 100 and 1k cycles. All

samples slid at 50.8 mm/s under a 250 N normal load. Wear track lengths for all samples were 41

mm, 31 mm and 19 mm. Wear rates were calculated as described above.

Page 52: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

52

Figure 3-1. Schematic of linear reciprocating tribometer with a flat-on-flat pin configuration, and

a six-axis load cell and several LVDTs for data acquisition.

Figure 3-2. A schematic of the stripe test run using a PTFE/5 wt. % α-alumina composite pin

against 304 stainless steel with a lapped finish. After each test, the pin was massed,

and the track length of the following test was shortened in order to expose a series of

transfer films in various stages of development. Schematic is adapted from Harris

2015.49

Page 53: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

53

Figure 3-3. Stripe tests were performed on the etched surfaces of unfilled PTFE and a PTFE/α-

alumina composite. The running films were removed, and two further stripe tests

were performed on the inner, un-modified polymer surfaces.

Page 54: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

54

CHAPTER 4

RESULTS AND DISCUSSION

Friction and Wear

The wear and friction performance of the PTFE/α-alumina composite was consistent with

previous studies of similar composite materials.43,44,46–48,63,93 Over the first 10k cycles, a run-in

period of moderately high wear was observed followed by a decrease in wear rate over the next

100k cycles to less than 10-6 mm3/(N·m) (Figure 4-1a). The coefficient of friction also decreased

slightly over the course of the experiment (Figure 4-1b), remaining near µ~0.19.

Stylus Profilometry

Figure 4-2 illustrates the dramatic changes undergone by the transfer film as it develops.

The measured root mean squared roughness (Rrms) of the stainless steel substrate was between

140 and 190 nm. Under 10k cycles, the average transfer film height was on the order of the

roughness of the surface. The average height of the transfer film after 100k cycles was

approximately 200 nm, with a local Rrms around 300 nm. The 1 million cycle film rose

approximately 1,000 nm above the baseline, with a local Rrms around 610 nm. Film profiles

shown in Figure 4-2 display the changing character of the roughness throughout the development

of the film. The most significant topographical changes in the transfer film occurred between

100k and 1 million cycles as the PTFE composite filled in the countersample grooves caused by

lapping and polishing.

X-Ray Photoelectron Spectroscopy

XPS (Figure 4-3) spectra provided a chemical time lapse of the transfer film formation.

Fluoropolymer transfer was detected after one cycle of sliding as evidenced by a peak around

292 eV. First cycle transfer of unfilled PTFE has previously been measured by surface plasmon

resonance33 and Auger117 techniques. In lab air, stainless steel countersurfaces are always coated

Page 55: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

55

with adventitious hydrocarbon contaminants and mixed oxides. A sputter depth profile was

performed to determine the approximate thickness of the contaminant layer and to measure the

composition of the bulk metal. After less than five minutes of sputtering, carbon and oxygen

compositions dropped below 10%. The contamination thickness was therefore likely less than 10

nm. At the end of the profile, the composition was 71% iron, 16% chromium, and 5% nickel -

consistent with bulk stainless steel given measurement uncertainties. In the XPS spectra of the

carbon region (shown in Figure 4-3a), the CF2 peak near 292 eV grew and the C-C/ C-H peak

near 285 eV shrank with as sliding cycles increased. This suggested that fluoropolymer coverage

increased within the transfer film region as sliding progressed. It also suggests that the surface is

cleaned of contaminants by the sliding composite, possibly due to the hard alumina filler

particles. This is further supported by observations of metals and oxides on the wear surface of

the polymer pin.

Figure 4-3b illustrates the changes in atomic concentration data from XPS experiments

on the transfer film as a function of sliding cycles. Transfer in the first cycle was succeeded by a

fluctuating fluoropolymer signal as debris was generated and ejected from the wear track as

wear. The concentrations of metal species from the steel (Fe, Cr) gradually decrease as the

transfer film thickens and the fluorine concentration rises. Alumina is first detected at 100 cycles

and reaches concentrations comparable to those in the bulk composite in the next few thousand

cycles. The wear rate drops to a steady state between 10k and 100k cycles as the protective

tribofilms grow. The small CF2 peak observed by XPS for the “0 cycle” experiment is probably

due to contamination from fluorocarbon wear debris on the surface of the steel counter sample

outside of the worn region.

Page 56: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

56

Infrared Spectroscopy

Fourier transform infrared (FTIR) spectroscopy performed within the exposed areas of

transfer film formed during the 1, 100k and 1M cycle tests, and on the cumulative wear debris

allowed for detailed chemical analysis of the evolution of the tribofilms and therefore of the wear

system. The spectral analysis is summarized in Figure 4-4. Spectra collected within the one cycle

transfer film area verified that the fluoropolymer had been immediately transferred to the metal.

However, the spectra also revealed an unusual set of peaks in the C-F region consisting of the

typical PTFE peaks at 1203 and 1149 cm-1 along with a new peak at 1253 cm-1 (Figure 4-4a).

This additional peak was originally derived by Moynihan in 1959 from first principles

calculations44 but has rarely been observed experimentally. It has previously been observed in

analysis of a PTFE powder film,118 for a PTFE sample run against a polyethylene film,119 and for

a PTFE sample run against 304 stainless steel.120 Lauer attributed some variations in intensity of

the peak near 1250 cm-1 to a stretching mode of the polymer molecule. He noted that the

intensity varied with the alignment of helical PTFE chains relative to the detector when using a

polarizer.120 This suggested that aligned PTFE chains were transferred to steel after a single

sliding pass, which agrees with surface plasmon resonance results from Krick et al.33, XPS

results from Uçar121, and the current study.

The extremely high molecular weight (~ tens of millions g/mol27) of PTFE means that the

granular polymer is very unlikely to transfer an entire chain to the metal surface due to the high

entanglement of the chains. Thus, for transfer to occur on a small scale prior to large scale

delamination, scission of the C-C backbone must occur. A mathematical argument can be made

to support the bond breaking and transfer of PTFE chains in the first cycle of sliding by

considering the intramolecular forces between the polymer chains and the metal countersurface.

PTFE fibrils have been shown to quickly and preferentially align in the direction of

Page 57: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

57

sliding.8,13,122–125 Thus, the model proposed is based on a balance between the sum of all the van

der Waals attractions between an aligned PTFE fibril and the metal surface, and the force

required to break an aligned PTFE fibril (Figure 4-5). Values for the model were chosen

conservatively so as to make it as unlikely as possible that fibrils would break rather than slide

across the surface.

The Hamaker solution for the attractive energy between a flat surface and a cylinder is a

function of the radius of the fibril, R, the length of the cylinder (or here, the fibril), L, the

separation distance between the fibril and the surface, d, and the Hamaker constant, A12 (the

theoretical value of the Hamaker constant for a PTFE-silica interaction (7.6 × 10−20 J) is used

here126). Expanded PTFE, in which PTFE filaments are highly aligned, exhibits a fibril tensile

strength of approximately 400-700 MPa. The energy equation is differentiated to yield the

attractive force between the fibril and the countersurface, Fadh, per unit length (Eq 4-1). This

attractive force, Fadh, multiplied by the friction coefficient of PTFE (~0.1 – a lower bound

considering the friction coefficient of the composite is nearly twice as high) equals the tensile

force applied to an aligned fibril in contact over the length L during sliding. This force is then set

equal to the ultimate tensile strength of the fibril, σf. Eq 4-2 allows us to solve for a critical fibril

length Lc (Eq 4-3). The interpretation is that any fibril in contact with the countersurface over a

length Lc or greater may be broken in sliding. This hypothesis is supported in the literature by

Makinson in 19645 and Brainard in 1973117. Additionally, evidence has previously been

published for the transfer of oriented films of PTFE onto a glass (silica) substrate during sliding

contact.127 It is evident that forces on aligned fibrils due to adhesion and friction are more than

sufficient (at entirely reasonable values of Lc) to break fibrils, and therefore C-C bonds at the

sliding surface.

Page 58: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

58

𝐹𝑎𝑑ℎ

𝐿=

𝐴12√𝑅

8√2∙𝑑52

(4-1)

µ𝐹𝑎𝑑ℎ = 𝜎𝑓 ∙ 𝜋𝑅2 (4-2)

𝐿𝑐~36𝜎𝑓

µ𝐴12∙ 𝑑

5

2 ∙ 𝑅3

2 (4-3)

In further support of instant transfer of PTFE, a control spectrum was obtained from the

an inner surface of the PTFE composite – accessed by slicing the sample with a razor blade -

with no sliding history, but the same thermal and environmental history. When the diamond ATR

crystal was held in place against the nascent surface, the spectrum contained the usual IR peaks

for bulk PTFE (Figure 4-5c i) at 1203 and 1149 cm-1. Once the diamond ATR crystal was

removed from contact with the sample, an IR spectrum was re-acquired in air before the crystal

was cleaned. Peaks in the C-F region were still visible, and in addition the spectrum included the

additional absorbance peak at 1253 cm-1 (Figure 4-5c ii) - identical to the one from the single

cycle sliding experiment (Figure 4-5c iii). PTFE chains must therefore have transferred from the

surface of the composite to the ATR crystal even after arguably static contact. The same three

peaks were also obtained from similar residual analysis of the ATR crystal after static contact

with an as-molded PTFE surface, which eliminated the razor cut as the source of the transferred

polymer chains. The transfer of PTFE to metal surfaces after static contact in high vacuum has

thus far only been observed by Auger photoelectron spectroscopy (AES).117

The IR spectra collected within the 100k cycle region or transfer film development

revealed new, broad peaks at 3388, 1650 and 1432 cm-1 in addition to the previously observed

PTFE backbone peaks at 1203 and 1149 cm-1. Within the 1M cycle region, these new peaks

dominate the spectrum (Figure 4-4b). Carboxylic acid end groups are often observed in

perfluoropolymers128. Their associated peaks are relatively sharp and are assigned to a mixture of

Page 59: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

59

both monomer (1775 cm-1) and dimer (1813 cm-1) forms. However, similar, much broader and

lower frequency carbonyl peaks at 1655 and 1441 cm-1 have been reported for instances where

fluorinated carboxylic acids chelated to metals as shown by Kajdas et al.129,130 In Kajdas’ study,

perfluorooctanoic acid (C7F15COOH) was coated onto a steel surface and then heated. A

reproduced spectrum is provided in Figure 4-6 a. Figure 4-6 c illustrates the similarity between

the IR spectrum from Kajdas’ chelated perfluoro- acid and the IR spectrum taken within the 1M

cycle film region, which suggests that the species present are identical, and that the broken PTFE

chains within the transfer film must be adhered to the metal surface in the same manner.

In a separate experiment at DuPont, a small molecule model compound (C6F13COOH)

and the same α-alumina were pre-mixed and dispersed in PTFE 7C for analysis by transmission

IR (Figure 4-6b). The resulting spectrum is very similar to that of the ATR-IR spectrum of the

running film (Figure 4-6d), which is in turn quite similar to that of the transfer film, suggesting

similar chemical changes occur in all. A control experiment of similar design without dispersed

α-alumina produced a spectrum that instead displayed the sharper monomer and dimer Rf-COOH

acid peaks at 1813 and 1775 cm-1, respectively. It is therefore suggested that the carboxylate

ends of PTFE chain ends chelate not only to the steel surface under the transfer film, but also to

the surfaces of the alumina filler particles. In fact, hydrocarbon carboxylic acids are already

known to react with the amphoteric surface of alumina particles. 131 It is not unreasonable, then,

for perfluorinated carboxylic acids (much stronger Brønsted acids than their hydrocarbon

analogs132) to react with and chelate to the alumina surface, even in the absence of excess heat.

ATR-IR analysis (Figure 4-4c) of the cumulative wear debris formed the end of the 1M

cycle track (shown schematically in Figure 3-2a) revealed large absorbance peaks for the PTFE

backbone, smaller monomer and dimer carboxylic acids peaks, and small peaks corresponding to

Page 60: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

60

the chelated salts. The cumulative wear debris is composed of polymer fragments shed during

the wear process. Thus, the chemical intermediates (carboxylic acids) were observed here in

addition to chelated salts (to the alumina particles) because further chemical modification or

chelation was not possible at all of the reaction sites after ejection from the wear track.

The results of the comparative IR analysis of the formation and evolution of the

tribofilms formed in this ultra-low wear polymer composite system led to the chemical

mechanism proposed in Figure 4-5. Figure 4-7 a & b illustrate the first step of the process: the

mechanochemical scission of the carbon-carbon bond in PTFE5,117 , which forms reactive

perfluoroalkyl radicals. The following steps of the mechanism (Figure 4-7 b-e) are identical to

the process of e-beam irradiation of high molecular weight PTFE in ambient air.14,103,133 The

perfluoroalkyl radicals then react with atmospheric oxygen to form a peroxy radical (Figure 4-

7c), which quickly decompose into more stable acyl fluoride end groups (Figure 4-7d). Acyl

fluoride end groups are unstable toward water and therefore hydrolyze in ambient humidity to

form carboxylic acids (Figure 4-7e). The dependence of the wear rate of this system on humidity

and vacuum environments has been described in previous chapters.43,46 HF is produced128 during

these steps, and likely goes on to form metal fluorides at the surface of the countersample. The

carboxylic acids chelate to the steel countersurface (Figure 4-7f) as evidenced by the IR studies

described. This strongly adheres them to the metal surface to form the robust transfer film that is

so closely associated with the ultra-low wear behavior of the composite.44,45

Jintang and Hongxin134 presented a similar mechanism for the mechanochemistry of

PTFE at the sliding interface, but do not mention the critical carboxylate groups which are the

source of the adhesion between the transfer film and the metal countersurface. The chelation of

carboxylate ended PTFE chains to the alumina fillers at the interface also reinforces the running

Page 61: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

61

surface of the polymer47,92, which increases the wear resistance of the polymer, and likely

reduces creep in the system. The tribochemical interaction between the PTFE/α-alumina

composite that breaks the C-C bonds is the first step in a complex cascade of events. Low wear

in this instance appears to be a property of the system created during sliding, and is heavily

dependent on the environment. Although conditions at the sliding interface are much milder than

what is required for thermal bond cleavage in PTFE,135 similar reaction products are detected in

both the running and transfer films at the interface (Figure 4-6). At low sliding speed, low

nominal contact pressure, and frictional heating not more than ~1° C removed from ambient

temperature, it is the coupling of mechanical and chemical effects that facilitates the formation of

the interfacial tribofilms necessary to maintain low wear over hundreds of thousands and even

millions of sliding cycles.

A study published by Khare around the same time as these results further investigated the

effects of environmental composition and counterface temperature on the chemical and

morphological response of the composite and its associated tribofilms.136 Khare et al. found that

across a range of humidity and oxygen contents, the composite always ran in to a low-wear

value, but in the case of the dry environments the wear of the composite increased again after

run-in. At high temperatures, wear was lower, the transfer film was thinner, and the carboxylate

signal was reduced. It is likely that increased temperature drove the reactions kinetically while

being insufficient to remove all of the necessary intermediate species at the interface (water,

oxygen). A wear experiment of the same type as many previously mentioned (flat on flat,

reciprocating over 25.4 mm at 50.8 mm/s at around 6 MPa) submerged the sliding interface of

the PTFE/α-alumina composite in question in distilled water. The sample ran for ten thousand

cycles at a wear rate nearly as high as unfilled PTFE, and three orders of magnitude higher than

Page 62: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

62

the dry composite after the same number of cycles (Figure 4-8). A transfer film never formed,

visible adhesion of debris was minimal to nonexistent and no color change to reddish brown was

observed. The reactions necessary for stable film formation could not proceed in the submerged

environment, and thus the wear of the sample remained very high

Khare et al. observed the formation of carboxylate salts in both high wear and low wear

systems, and found no correlation between the oxygen content of the test chamber and the steady

state wear rate of the composite. However, they could not rule out the presence of trace amounts

of oxygen that could still be sufficient for the mechanism to proceed as suggested here.

Furthermore, their IR studies assigned peaks at 1315 and 1360 cm-1 to shortened chain lengths

within the polymer, supporting the hypothesis of mechanical chain scission described above, and

interface pairing experiments showed (in agreement with Bahadur and Tabor36) that an existing

transfer film did not reduce the wear of a fresh composite pin, and that an existing running film

significantly reduced the run-in volume against a fresh countersurface. The reinforcement of the

composite due to the chelation of PTFE chains to the alumina filler particles must then be the

dominant source of wear abatement, and the transfer films are an effect thereof.

It was then supposed that if a precursor to the running film could be formed at on the

sliding surface of the composite that the run in period of high wear might be abated or avoided

entirely. Samples of unfilled PTFE and of the PTFE/α-alumina composite were etched and tested

as described previously. Wear results from the stripe tests are given in Figure 4-9. The etched

surface of the unfilled PTFE sample lowered the wear rate by almost two orders of magnitude

over the first ten cycles (~0.5 m) of sliding, but quickly wore through over the next hundred

meters. The wear rate then returned to a higher value, closer to that of the unmodified PTFE

sample. The wear rate of the etched composite over the first 1000 cycles (~50 m) was reduced by

Page 63: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

63

nearly an order of magnitude compared to the inner, unmodified surface, and over the next 10k

cycles, the etched sample was observed to have lost no mass within the error of the balance. The

final, steady state wear rate of both the etched and unmodified composite surfaces was very

similar, but the etching treatment did indeed appear to greatly reduce the run-in volume during

the beginning of the test.

The results of these wear tests strongly correlate with the appearance of the transfer and

running films formed (Figure 4-10). The highest wearing sample, the unfilled, unmodified PTFE,

left large wear platelets in the patchy transfer pattern characteristic of PTFE sliding at more than

~10mm/s.5 The etched surface of the unfilled PTFE performed slightly better, and the debris

within the wear track resembles the brown transfer film characteristic of the composite at the

metal surface, but is covered with PTFE wear platelets that formed once the etched film was

worn through and the polymer resumed increased wear. The edges of the etched surface not on

the leading or trailing faces were not worn through, but this is possibly due to a leveling issue

between the polymer pin and the metal countersurface. The wear results should not be

compromised however, as the wear rate is normalized by the supported load and such the wear

area is reduced but the nominal contact pressure is increased. The difference in contact pressure

should not be enough to invoke any significant dependence of wear rate on load that may exist.

The unmodified inner surface of the PTFE/α-alumina composite performed as expected

and formed the characteristic brown transfer film. The appearance of the running film changed

very little over the duration of the stripe test. Most notably, the etched surface of the composite

formed a much thinner transfer film, as supported by the markedly lower run-in wear volume.

Again, the running surface of the polymer hardly changed in appearance throughout the duration

of the stripe test, but based on the presence of any visible transfer film it is possible that the

Page 64: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

64

etched surface layer may have been mostly or completely replaced by a running film formed in-

situ. It is likely that the carboxylic acid end groups in the etched layer will have immediately

chelated to the alumina particles at the near surface. During the run-in period, the polymer is

reinforced as the large, friable alumina particles break up, further increasing the concentration of

virtual cross-links in the running film. Transfer to the countersurface is minimal during this

period, but a thin layer of chelated PTFE chains still appears to fill in the negative features of the

steel and provide a softer, reinforced polymer surface for the polymer pin to slide against. The

results of these etching experiments support Khare’s observation that a pre-existing running film

greatly reduces run-in of the composite, and confirm that the introduction of carboxylic acid

chain ends at the sliding interface speed up the transition to ultra-low wear.

Figure 4-1. Wear rates and friction coefficients for each experiment run as a part of the stripe

test. The wear rate of the composite behaved as expected and ran in to ultra-low wear

despite the alteration of the wear track length (a), and decreased through a run-in and

transition period to a steady state on the order of 10-7 mm3/N∙m. The friction

coefficient (b) increased over the first hundred cycles and then ran in to around 0.17

over the million cycles of sliding.

Page 65: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

65

Figure 4-2. Stylus profiles of transfer films after 1 cycle, 100k cycles, and 1M cycles of

development. Most transfer film growth took place between 100k and 1M cycles.

Each line scan provided is an average of 50 line scans at 0.04 mm intervals.

Figure 4-3. High resolution XPS of the transfer film in various stages of development

demonstrate some of the chemical changes present in the films, and the evolution of

atomic concentration as film formation progresses. a) C1s spectra from exposed areas

of transfer film after 0, 1, 100k and 1M sliding cycles and b) atomic concentrations in

regions of transfer film development after 0, 1, 100k and 1M sliding cycles compared

to bare stainless steel.

Page 66: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

66

Figure 4-4. Infrared reflectance results from the transfer film on a stainless steel surface after (a)

one cycle of sliding, (b) 100k and 1M cycles, and (c) ATR-IR spectrum of cumulative

wear debris. (Reproduced from Harris 201549)

Page 67: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

67

Figure 4-5. A Hamaker solution for the attractive energy between a flat surface and a cylinder is

used to support the hypothesis that PTFE fibrils break and transfer during the first

cycle of sliding. (a) A PTFE fibril of average radius R and separation distance d from

a countersample contacts the sliding countersurface over a length L. The fibril

experiences an adhesive force due to van der Waals interactions at the surface (b) A

plot of the critical fibril length Lc as it changes with average fibril radius R for

separations d=1, 2 and 10 Å demonstrates the increase in Lc as R and d increase.

Shaded circles are in place to represent fibrils of various aspect ratios. (c) (i) IR

spectra of bulk PTFE, (ii) the residue spectrum after pulling away from the bulk

PTFE, and (iii) the 1 cycle transfer film. (Adapted from Harris 201549).

Page 68: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

68

Figure 4-6. The carbonyl region of IR spectra of the tribofilms is compared to small molecule

model reactions with perfluorinated carboxylic acids: (a) the IR spectrum

published by Kajdas and Przedlacki130 ascribed to the chelated salt of

perfluorooctanoic acid on a steel countersurface. (b) the IR spectrum collected

from a pressed PTFE film filled with a dispersion of alumina particles and

perfluoroheptanoic acid (c) the IR reflectance spectrum acquired within the 1M

cycle transfer film, and (d) ATR-IR spectrum of the running surface of the

polymer composite pin after 1M cycles. (Adapted from Harris 201549)

Page 69: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

69

Figure 4-7. The chemical mechanism responsible for PTFE/α-alumina tribofilm formation and

adhesion. Chain scission of PTFE (a) is caused during sliding due to frictional forces

and adhesive forces between the PTFE fibrils and the metal countersurface (b).

Environmental oxygen reacts with the radicals created (c). The unstable end groups

decompose to form acyl fluorides (d). Moisture in the air hydrolyzes these groups,

forming carboxylic acids (e). Carboxylic acid end groups react with and chelate to the

surface of the metal and alumina particles (f). Figure adapted from Harris 2015.49

Page 70: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

70

Figure 4-8. The wear rate of the PTFE/α-alumina composite and neat PTFE compared to that of

the submerged composite. The wear rate of the composite was three orders of

magnitude higher against 304 stainless steel when submerged in distilled water than

against the same type of countersurface in laboratory air.

Figure 4-9. The wear rates of the etched and unmodified surfaces of unfilled PTFE and of the

PTFE/α-alumina composite are plotted versus the total sliding distance. Sliding

distance is used rather than the number of sliding cycles because of the variable

length of the wear track during a stripe test. The mass loss of the PTFE/α-alumina

composite after the second (10k cycle) test was measured to be zero, which cannot be

represented on a log-log plot and is therefore omitted from the figure.

Page 71: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

71

Figure 4-10. Optical images of the transfer and running films of the etched and unmodified

surfaces of unfilled PTFE and of the PTFE/α-alumina composite. Transfer film

images are displayed alongside images of the running films of each of the samples

after the end of each stripe test (10k cycles total for the unfilled PTFE samples, and

100k cycles total for the composite samples). Photos taken by author in Spring 2016.

Page 72: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

72

CHAPTER 5

CONCLUSIONS

Transfer of PTFE to the metal countersurface occurs after a single cycle of sliding, and

even after static contact, as evidenced by FTIR (Figure 4-4). The polymer/alumina composite

then gently abrades/cleanses the adventitious carbon and surface oxides from the surface of the

steel as it slides. Without removal of this contaminant layer and native oxides, the chelation of

the carboxyl terminated fluoropolymer would be much less likely, which would limit the

adhesion of the tribofilm. The hypothesis of mechanochemical chain scission is supported by a

force balance model based on van der Waals interactions between the polymer and the metal

countersurface, described in Figure 4-5. The wear rate of the composite was high during the run-

in period and did not fall below 10-6 mm3/(N·m) until the 100k cycle test. The change in wear

rate from high to low coincided with the appearance of IR peaks at 1650 and 1432 cm-1 in the

transfer and running film spectra (Figure 4-4b, Figure 4-6c and d) which are indicative of the

presence of chelated carboxylate polymer chain ends (Figure 4-7f). The concentration of

chelated salts increased significantly during the 1M cycle test (Figure 4-4b) as more and more

chains broke and reacted with the metal and Al2O3 surfaces. Carboxylic acid chain ends that did

not chelate to a surface were ejected as wear debris, and were the source of the carboxylic acid

monomers and dimers identified by IR of the cumulative wear debris (Figure 4-4c).

The formation of an ultralow wear PTFE transfer film on 304 stainless steel involves the

chemical interaction between the polymer composite, the embedded alumina particles, the

ambient atmosphere, and the metal countersurface. This is a cycle-dependent process (under

reciprocating conditions) that relies on the mechanical input of energy to cause chain scission,

which in turn initiates the radical-driven mechanism of transfer film formation. The reactions

proposed in this mechanism have been observed previously - in non-tribological settings – but in

Page 73: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

73

this context they neatly describe the chemistry behind the observed wear behavior of this

PTFE/α-alumina composite. Figure 5-1 summarizes this ultralow wear system, which arises from

a complex set of variables that together allow chemical modification of both sliding surfaces

under relatively mild conditions. The ultra-low wear behavior of the PTFE/α-alumina composite

described is driven by reactions within the running film, and is accompanied by similar reactions

with the countersurface, though the latter are an effect rather than a cause of the low wear

behavior. High wear composites remove debris quickly from contact, but in reinforced, ultra-low

wear systems, debris is retained and may react with species present.

It is shown in this work that PTFE, though widely regarded as an inert, environmentally

insensitive polymer is in fact quite reactive within the sliding contact. An explanation for chain

scission of PTFE under mild conditions has introduced a new perspective regarding the behavior

of ultra-low wear PTFE/α-alumina composites. The following steps of the mechanism have been

understood in the context of fluoropolymers, but were never applied to the wear process because

the presence of the initial carbon radical due to the scission of the carbon backbone had not been

considered. The performance of these ultra-low wear composites cannot be attributed to any

single component of the system, but the chemical bonds formed during sliding between the

polymer chains, the alumina particles and the countersurface are crucial to the formation and

longevity of the tribofilms. This discovery has not only changed the industry’s understanding of

these composites, but introduces the possibility of simple surface modifications to the

composites themselves rather than to their countersurfaces that further increase their usable

lifetime.

Page 74: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

74

Figure 5-1. Radical chemistry at the sliding interface proceeds despite mild conditions (low

speed, low nominal contact pressure, and low frictional temperature change). The

circulation, rather than ejection, of debris between the transfer and running films is

likely key to the high cycle maintenance of ultralow wear. Adapted from Harris

2015.49

Page 75: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

75

APPENDIX

FUTURE CONSIDERATIONS: COUNTERFACE EFFECTS

Wear and Friction Experiments on Additional Countersurfaces

Throughout the studies on the relationship between PTFE, the α-alumina filler and the

stainless steel countersurface, which has become the standard for polymer wear testing, many

wear and friction studies were performed on alternate metal countersurfaces in order to elucidate

more characteristics of the polymer/metal system. Wear and friction results for a suite of

experiments spanning multiple years is presented in Figure A-1 and Figure A-2. Experiments

were all performed using the same polymer pin dimensions, sliding speeds and track lengths as

usual (6.3x6.3 mm polymer running surface, 250 N normal load, 50.8 mm/s on a 25.4 mm long

wear track). The error in measured wear rate is left out of the plots due to the logarithmic scale

and density of data, but in the case of tests run out to 100,000 cycles or more, the error in

measured wear rates during steady state is frequently, if not always very small, and usually an

order of magnitude less than the measured wear rate.

Against most steels, which were lapped and polished to average roughnesses around 150

nm, the measured wear of the composite behaved as expected, with a period of run-in followed

by steady state low wear between 10-8 and 10-5 mm3/N∙m, as illustrated in Figure A-1. The

friction coefficients of the polymer/steel systems were measured to be mostly around 0.2.

Strangely, in the case of the high strength 4340 countersample, the friction coefficient increased

dramatically after 10,000 cycles, but the wear rate of the composite remained at a lot, steady

state value. The cause of this behavior is unknown, but it seems to highlight the robust nature of

the tribofilms formed, as the wear performance was not compromised.

Three copper countersurfaces, pure copper (CDA110), beryllium copper (CDA 172) and

naval brass (C464) were selected for testing due to their wide range of mechanical properties and

Page 76: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

76

because copper was perceived as a fairly reactive metal that could possibly facilitate reactions at

the interface, which could increase the chance of stable film formation. Metal surfaces were

lapped and polished to ~150 nm Ra. The wear results are summarized in Figure A-1. Both the

pure copper and the naval brass performed relatively poorly, with composite wear rates in the

range of 10-5, while the beryllium copper rivalled the performance of 304 stainless steel. The

friction was highest during the beryllium copper test, and especially low against pure copper. A

wear summary of the copper tests is provided in Figure A-3 in conjunction with friction traces

for each, and photos of the polymer and metal wear surfaces. On the poor performing pure

copper surface, the polymer film is patchy and a greenish tint at the surface suggests the presence

of copper carbonates. Furthermore, the wear debris at the track ends also appears to be filled with

green copper carbonate and black copper oxides, and the countersurface is abraded. The running

film appears poorly developed and likely never completely formed under gross wear conditions.

The beryllium copper films were well formed, and the naval brass countersurface was abraded

and accompanied by a good deal of black wear debris, likely degraded PTFE and copper oxide.

The friction trace of the beryllium copper test also includes drops from around 0.3 to around 0.2

at cycle numbers corresponding to mass measurements, for which the polymer was removed

from contact with the transfer film. It is speculated that this may be due to water adsorption

occurring at the film surfaces during these periods.

Friction and wear tests were also performed against 24k gold, pure platinum, and pure

lead. The platinum and gold surfaces were smooth, mirrored surfaces when they arrived and

were not polished further. The lead countersample was polished by hand to around 400 nm Ra.

Wear and friction results are given in Figures A-1 and A-2 respectively. The gold countersurface

itself failed before reaching 10,000 cycles. Because of counterface failure, the later wear rate

Page 77: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

77

includes mass uptake from gold transfer, and the later friction coefficient includes the ploughing

force to yield the gold. A more detailed summary of the experiment is provided by Figure A-4.

No transfer film could be observed on the countersurface after its failure, and a great deal of gold

transferred to the polymer running surface, visible in the optical images in Figure A-4b. Most

interestingly, the polymer running surface not obscured by significant gold transfer appeared

bluish in tint. It is possible to create visibly blue solution of large (~100nm) gold nanoparticles,

or by adding excess salt to a solution of smaller gold nanoparticles, and a blue solution of gold

‘nano-urchins’ is also possible137. Chemical interaction between the composite components and

the gold countersurface was not expected. The platinum countersurface held up during the wear

test, although the ultimate performance of the composite was just shy of very low wear, never

dipping into the 10-7 mm3/N∙m range. Increases in friction to 0.2 at mass measurement locations

are visible in the platinum friction and are again attributed to water adsorption at the films. The

lead countersurface held up surprisingly well for around 150,000 cycles. The wear of the

polymer during this period was also fairly low, but not as low as is usually observed against

stainless steel. The transfer film that formed before the failure of the lead countersurface

appeared to be extremely thin – thin enough to appear iridescent. The wear track before and after

failure is given in Figure A-4b. The polymer surface from the lead test was also discolored as

during the gold test, but the color observed was yellow. The yellow tint persisted after the failure

of the countersurface during which the polymer pin was also misshaped and subject to lead

transfer. Both platinum and lead were thought to be more reactive than the gold, and therefore

were considered more likely candidates for stable film formation.

Three aluminum alloys (6061 T6, 7075 T6, 1100) were chosen as countersurfaces for the

PTFE composite as a negative control, based on the classically poor tribological performance of

Page 78: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

78

aluminum surfaces which is normally abated using a surface coating, particulate reinforcement

of the metal, or anodizing138–140. Surprisingly, all three of the alloys outperformed the stainless

steel countersurface used for reference with respect to the wear of the composite (Figure A-1) at

a slightly higher coefficient of friction (Figure A-2). Traces of the friction coefficient of the

composite against the three aluminum alloys and against stainless steel are provided in Figure A-

5, along with optical images of the transfer films. The films in each case appeared robust and

uniform, and were all gray or white in color.

Infrared Spectroscopy and X-Ray Analysis of Transfer and Running Films

The transfer films formed on C1018 and M2, the worst and best performing steels,

respectively, were analyzed using FT-IR in reflectance mode in the same manner as the 304

stainless countersurfaces described in previous chapters. A comparison of the resulting spectra is

given in Figure A-6 with arbitrary intensity. All spectra taken from the wear tracks against steels

show C-F stretch near 1200 cm-1 and carboxylate salt C=O peaks near 1434 and 1655 cm-1 on

C1018, and 1430 and 1670 cm-1 on M2. Water is indicated by a broad O-H peak near 3300 cm-1.

These are likely to indicate hydrated perfluoro carboxylate salt material(s). A broad peak below

900 cm-1 is likely due to metal-oxide M-O stretch, probably from aluminum oxide. Minor peaks

near 949, 993, and 1311 cm-1 in the M2 spectrum are unassigned at this time. It is important to

note that the presence of chelated polymer chains at the metal surface does not in itself predict

low wear of the polymer composite. The polymer chains may chelate not only to the metal

surface, but also to the filler particles, and chelation is observed in the wear debris as well as in

the transfer and running films.

Representative IR spectra of the transfer films on CDA 110 and CDA 172 are provided

(with arbitrary intensity) in Figure A-7, again in comparison to the stainless steel transfer film.

Both copper spectra show C-F stretch near 1200 cm-1 and carboxylate salt C=O peaks. The

Page 79: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

79

higher frequency C=O peak on pure copper show a sharp location near 1615 cm-1 in most spectra

and a broader location in the 1650-1660 cm-1 range. The sharper peak could be from the salt and

the broader peak from water of hydration. The second carboxylate peak is near 1435 cm-1.

Carboxylate peaks on the beryllium copper countersurface are seen near 1435 and 1655 cm-1.

Water is indicated in both spectra by a broad O-H peak near 3400 cm-1, consistent with hydrated

perfluoro carboxylate salt materials. Again, a broad peak below 900 cm-1 is likely due to metal-

oxide M-O stretch, probably from aluminum oxide. A sharp peak near 3643 cm-1 in most spectra

may be from the water of hydration of an inorganic component. A broad peak below 900 cm-1 is

likely due to metal-oxide M-O stretch, probably from aluminum oxide. Minor peaks near 1320

and 1363 cm-1 are consistent and are likely due to the salt and may have to do with an oxalate-

like structure but are not assigned at this time. Again, the presence of carboxylate salts suggests

chain scission in PTFE and reactivity at the scission sites, but is not a predictor of low wear.

XPS spectra (C1s) were collected from the running films created during the

lead/platinum/gold tests, and are given with arbitrary intensity in Figure A-8 in comparison to

that of nascent PTFE and that of a running film formed against 304 stainless steel. In all cases,

chain scission and end group chemistry appear to be present. However, precise bonding

information regarding chelation is not available from this analysis, and IR spectra are not yet

available.

In the case of all three transfer films on the aluminum alloys, C-F stretch near 1200 cm-1,

water (the broad O-H peak near 3250 cm-1) and carboxylate salt peaks (1425 and 1672 cm-1 in

the case of 6061, 1425 and 1655 in the case of 7075, and 1417 and 1657 cm-1 in the case of

1100) are present in the FT-IR spectra, shown with arbitrary intensity in Figure A-9. As is the

Page 80: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

80

case with most of the other transfer films, these results are consistent with the presence of a

hydrated perfluoro carboxylate salt material.

Surface and Sub-Surface Evolution of the Aluminum Countersamples

Further analysis of the aluminum transfer films and countersamples was performed

because of the remarkable nature of the wear results, and results were considered in comparison

to the case of 304 stainless steel. Against each of the three aluminum countersurfaces, a thin,

stable transfer film had formed, the polymer composite reached ultra-low wear, and contrary to

expectation the aluminum countersamples were not subjected to delamination or other visible

mechanical failure. Stylus profilometry of the transfer films shows more variability across the

aluminum samples than against stainless steel (Figure A-10), and it appears that in the case of the

6061 T6 countersurface, the polymer composite wore into the metal surface as the low wear self-

mated film system formed.

The interaction between the polymer composite and the aluminum countersurfaces was

then assumed to extend into the subsurface of the metal. Removal of the polymer film allows

inspection of the changes undergone by the metal surfaces during testing. The transfer films

formed by polymer composites against stainless steel are frequently easily removed using boiling

water, and the same process was applied to two of the aluminum countersurfaces (6061 T6 and

Al 7075 T6). IR spectra taken in various areas within the wear regions after boiling water

treatment are provided for 304 stainless steel, Al 6061 T6, and Al 7075 T6 in Figure A-11.

Within the direction change region (characteristic of reciprocating tests) and the center of the

wear track of the stainless steel countersample, no significant peaks were detected, indicating no

residual fluorinated material on the metal surface.

A spectrum taken within the striated wear track area of the 6061 T6 countersurface

showed no fluorinated material, but spectra collected in the direction-change area and the end of

Page 81: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

81

the travel adjacent to the direction change region revealed the persisting presence of a hydrated

perfluoro carboxylate salt. A peak near 1049 cm-1 in the direction change area is visible, and may

have to do with aluminum chemistry. A spectrum from the striated wear track area of the 7075

T6 countersurface provides weak evidence for fluorinated material, probably the carboxylate

salt. The end-region spectrum taken at the end of the wear track area adjacent to the direction

change region shows no significant fluorine presence. A spectrum taken within the 7075 T6

wear track includes a new peak near 1065 cm-1 that may have to do with aluminum chemistry. A

peak at 1065 cm-1 is associated in vaccine science with a structural hydroxide environment of

poorly crystalline boehmite, an aluminum oxide hydroxide mineral that may have to do with the

boiling treatment.141 A spectrum from the direction-change region of the 7075 T6 countersurface

clearly indicates the persisting presence of a hydrated perfluoro carboxylate salt, and includes the

1065 cm-1 peak. The C-F peaks near 1200 cm-1 were more consistent in frequency compared to

the 6061 T6 countersurface. The O-H region above 3000 cm-1 is variable in intensity for both

aluminum countersurfaces analyzed, and may have to do with hydration of aluminum, hydroxide

formation and/or hydrated carboxylate salt when present.

In order to further investigate the interaction of the composite and the transfer film with

the aluminum countersurface, a Focused Ion Beam (FIB) microscope was used to cut trenches

within and outside of the wear track. Images acquired at the University of Florida are given in

Figure A-12a (within the wear track) and A-12b (in the nascent metal away from the wear track),

and images acquired later at the Center for Integrated Nanotechnology (CINT) in Albuquerque

are provided in Figure A-12c. Within nearly every FIB trench, distinct and unusual features are

visible in the region 1-2 µm beneath the surface. Grain size appears greatly reduced at the near

surface. In the trenches taken within the wear tracks, it is presumed that the black stripe at the

Page 82: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

82

surface below the protective platinum layer and the carbon coat is the polymer composite’s

transfer film. The large circular feature beneath the surface of the 1100 is of particular note, and

may suggest the presence of polymer penetration into the surface of the metal, which could

explain the persistence of fluorine signal after the removal of the visible transfer film with

boiling water. However, because the near surface appears affected within the nascent metals, the

features visible cannot be confidently attributed in full or in part to the sliding composite contact.

In the months following the FIB investigation, it was still thought that the polymer

composite was somehow uniquely suited for ultra-low wear sliding contact against aluminum

countersurfaces, and many further wear tests were conducted against aluminum 6061 T6

countersurfaces from various sources. Results of the wear tests are summarized in Figure A-13.

The original, high performance aluminum countersurfaces were delivered as ‘lapped’, at a

quoted average roughness of around 150 nm. Attempts in lab to replicate the success of the

original surface via wet polishing or lapping against a cast iron wheel with silicon carbide/oil

lapping compound produced only countersurfaces against which the wear of the composite was

orders of magnitude higher than against the original samples, seemingly regardless of measured

average surface roughness. It was apparent at this juncture that only the aluminum

countersurfaces prepared by the original supplier, Alabama Specialty Products Incorporated

(ALSPI), were consistently achieving ultra-low wear with the PTFE/α-alumina composite,

despite attempts to reproduce their polishing process exactly. Indeed, aluminum samples from

other suppliers that were sent to ALSPI to be lapped also held up in wear tests and saw the

polymer composite reach ultra-low wear conditions for hundreds of thousands of cycles.

It was then supposed that perhaps some component of their lapping process was unique to

their location. Samples of the ALSPI lapped 6061 T6 countersurfaces and of the poor-performing

Page 83: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

83

coupons from other sources and subjected to various surface treatments were shipped to the

University of Illinois at Urbana-Champaign for further FIB analysis for comparison, and for

nanoindentation as an initial check for any changes to mechanical properties of the surface due to

the ALSPI lapping process. The FIB images taken are shown in Figure A-14, and confirm a large

difference between the surfaces produced at ALSPI and the others tested. The images from the as

received, unpolished surfaces (samples 1 and 2) show no distinct features at the near surface, and

those of polished or milled (samples 4 and 5) show some grain refinement. However, the ALSPI

lapped countersurface (sample 3) again was seen to possess an unusually disrupted structure. It is

important to note that the features in the ALSPI lapped surfaces illustrated here are seen in a

large fraction of the trenches milled. Nanoindentation (Berkovich tip, 5 mN) was performed on

each countersurface and the hardness values and elastic moduli reported in Figure A-15 are the

average results from nine indents per countersurface. No significant difference in hardness or

modulus was found between the surfaces tested despite the large disparity in tribological

performance, although a brief Energy Dispersive X-Ray Spectroscopy (EDS) scan indicated the

presence of iron in the surface of the ALSPI sample (3), and not in sample 1.

An in-person visit to ALSPI allowed for closer inspection of the lapping process that was

producing such remarkable aluminum countersamples. The lapping wheel used was cast iron,

and coated with thick, oil lapping compound that was not cleaned or changed between different

metal samples. The aluminum samples were lapped for around ten minutes under only the dead

weight load of the sample holder, wiped clean with a dry rag, and polished briefly and gently by

hand using 600 grit silicon carbide sandpaper that was well used previously to smooth mostly

iron samples. Figure A-16 includes several images of the process. Samples were collected of the

lapping compound, which was later suspended in ethanol and filtered for solids to be analyzed

Page 84: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

84

via EDS. One resulting spectrum is provided in Figure A-17, and indicates the presence of not

only the original components of the lapping compound (Si, C, O), but an extremely large amount

of contamination by metals processed in the facility (Co, Cr, Fe, Al). One resulting hypothesis

was that the lapping process embedded iron from the lapping wheel (which wears over time, with

use) and other metal particles left in the lapping compound into the relatively soft and malleable

surface of the aluminum, acting as reinforcement, though nano-indentation revealed no hardness

or modulus effects.

Collaborators at DuPont collected EDS maps of the near surface in FIB trenches milled in

a 6061 T6 sample lapped at ALSPI, and in another subjected only to hand polishing on wet

silicon carbide sandpaper. Backscatter micrographs and the corresponding EDS maps in Figure

A-18 show that both samples contain iron (consistent with the composition of the alloy), but the

distribution thereof is quite different between them. In the hand polished sample, against which

the wear of the PTFE/α-alumina composite was quite high, the distribution of iron is regular and

random. In the ALSPI lapped sample, against which the wear of the composite was very low, the

iron is distributed throughout the sample, but is also concentrated in horizontal, parallel tracks

around 10 µm apart. It is presumed that this is due to the lapping process, but the mechanism is

as yet unknown. However, the nano-crystallization of the surface of aluminum and other

crystalline in a rubbing contact has been described for decades as the Beilby layer, the depth of

which is dependent on the ‘vigor’ of the polishing action.142–147 Recently, improvements to the

wear resistance of aluminum via mechanical surface modification were described by Nimura in

in the case of fretting in oil.148 It is possible that correctly replicating the surface modifications

seen in aluminum samples lapped by ALSPI, a simple, low cost method of preparing aluminum

Page 85: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

85

surfaces for dry tribological contact with polymer composites could be established, reducing the

need for coatings or more complex surface treatments that are commonly used today.

Figure A-1. A summary of the wear results from PTFE/α-alumina composite experiments against

various metal countersurfaces.

Page 86: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

86

Figure A-2. A summary of the friction results from PTFE/α-alumina composite experiments

against various metal countersurfaces.

Page 87: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

87

Figure A-3. A summary of the results of PTFE/α-alumina composite experiments against various

copper containing countersurfaces. a) The total wear rate as sliding progressed, b)

optical images of the wear tracks and running films and c) friction traces spanning the

length of each experiment. Photos taken by author in 2014.

Page 88: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

88

Figure A-4. A summary of the results of PTFE/α-alumina composite experiments against lead,

gold and platinum. a) The total wear rate as sliding progressed, b) optical images of

the wear tracks and running films and c) friction traces spanning the length of each

experiment. Photos taken by author in 2014.

Page 89: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

89

Figure A-5. Friction traces from the PTFE/α-alumina composite experiments against three

aluminum alloys and against stainless steel, and optical images of the transfer films

from the aluminum experiments. Photos taken by author in 2014.

Figure A-6. FT-IR spectra taken within the transfer films formed by the PTFE/α-alumina

composite against three types of steel are compared to the spectrum taken within a

transfer film formed on stainless steel.

Page 90: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

90

Figure A-7. FT-IR spectra taken within the transfer films formed by the PTFE/α-alumina

composite against three different copper alloys are compared to the spectrum taken

within a transfer film formed on stainless steel.

Figure A-8. XPS spectra taken from the running films formed by sliding the PTFE/α-alumina

composite against platinum, gold, lead, and stainless steel are compared to a spectrum

taken from the nascent polymer surface.

Page 91: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

91

Figure A-9. FT-IR spectra taken within the transfer films formed by the PTFE/α-alumina

composite against three aluminum alloys are compared to the spectrum taken within a

transfer film formed on stainless steel.

Figure A-10. Stylus profilometry traces taken across the center of transfer films formed by the

PTFE/α-alumina composite against three aluminum alloys are compared to the profile

taken across a transfer film formed on stainless steel.

Page 92: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

92

Figure A-11. FT-IR spectra illustrate the changes in surface chemistry before and after

attempting to remove the transfer films from 304 stainless steel, Al 6061 T6 and Al

7075 T6 with boiling water

Page 93: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

93

Figure A-12. FIB cross sections taken within the wear track and in nascent areas of the three

aluminum alloys tested. a) micrographs acquired within the wear tracks of all three

alloys, b) micrographs acquired in nascent areas of the three alloys, and c)

micrographs acquired within the wear track of an Al 6061 T6 countersurface. Images

in a & b were collected by Nick Rudowski at the University of Florida in Fall 2013.

Images in c were collected at CINT in Albuquerque, NM by the author in Spring

2014.

Page 94: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

94

Figure A-13. A summary of the wear results of the PTFE/α-alumina composite against a number

of Al 6061 T6 countersurfaces purchased from various locations and prepared in

different ways.

Page 95: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

95

Figure A-14. Micrographs taken within FIB trenches in the nascent surfaces of five differently

prepared Al 6061 T6 surfaces. Micrographs collected by Matthew Bresin in Spring

2015 at the University of Illinois at Urbana-Champaign.

Figure A-15. Hardness and modulus results from nanoindentations performed on the surfaces of

five differently prepared Al 6061 T6 surfaces. No significant difference is observed.

Page 96: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

96

Figure A-16. A series of images depicting the lapping and polishing process at ALSPI. Photos

taken by the author in Summer 2015 at Alabama Specialty Products Inc. in Munford,

AL.

Page 97: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

97

Figure A-17. An EDS spectrum collected from a sample of the solids separated from the lapping

compound taken directly from the lapping wheel used by ALSPI.

Page 98: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

98

Figure A-18. Backscattered electron micrographs and EDS maps of the subsurface (~1 mm

depth) of two Al 6061 T6 samples, one untreated, and one lapped and polished by

ALSPI. The images on the left were taken from a FIB trench in an untreated sample,

and the images on the right were taken from a FIB trench in a sample lapped and

polished by ALSPI.

Page 99: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

99

LIST OF REFERENCES

(1) Persson, B. N. J. Sliding friction: physical principles and applications; Springer, 2000;

Vol. 1.

(2) Shooter, K. V; Tabor, D. Proc. Phys. Soc. Sect. B 1952, 65 (9), 661–671.

(3) King, R. F.; Tabor, D. Proc. Phys. Soc. Sect. B 1953, 66 (9), 728–736.

(4) Flom, D. G.; Porile, N. T. J. Appl. Phys. 1955, 26 (9), 1088.

(5) Makinson, K. R.; Tabor, D. Proc. R. Soc. A Math. Phys. Eng. Sci. 1964, 281 (1384), 49–

61.

(6) Blanchet, T. A.; Kennedy, F. E. Wear 1992, 153 (1), 229–243.

(7) Deli, G.; Qunji, X.; Hongli, W. Wear 1989, 134 (2), 283–295.

(8) Tanaka, K.; Uchiyama, Y.; Toyooka, S. Wear 1973, 23 (2), 153–172.

(9) Tanaka, K.; Kawakami, S. Wear 1982, 79 (2), 221–234.

(10) Blanchet, T. A.; Kennedy, F. E.; Jayne, D. T. 1993.

(11) Deli, G.; Bing, Z.; Qun-Ji, X.; Hong-Li, W. Wear 1990, 137 (1), 25–39.

(12) Lai, S.-Q.; Yue, L.; Li, T.-S.; Hu, Z.-M. Wear 2006, 260 (4-5), 462–468.

(13) Wheele, D. R. Wear 1981, 66 (3), 355–365.

(14) Fischer, D.; Lappan, U.; Hopfe, I.; Eichhorn, K.-J.; Lunkwitz, K. Polymer (Guildf). 1998,

39 (3), 573–582.

(15) Yuan, X.-D.; Yang, X.-J. Wear 2010, 269 (3–4), 291–297.

(16) Fluoropolymers 2; Hougham, G., Cassidy, P. E., Johns, K., Davidson, T., Eds.; Topics in

Applied Chemistry; Kluwer Academic Publishers: Boston, 2002.

(17) Patrick, C.; Stacey, M.; Tatlow, J.; Sharpe, A. 1961.

(18) Renfrew, M. M.; Lewis, E. E. Ind. Eng. Chem. 1946, 38 (9), 870–877.

(19) Sperati, C. A.; Starkweather, H. W. Advances in Polymer Science; Advances in Polymer

Science; Springer-Verlag: Berlin/Heidelberg, 1961; Vol. 2/4.

(20) Plunkett, R. J. Tetrafluoroethylene polymers. US 2230654 A, February 4, 1941.

Page 100: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

100

(21) Hintzer, K.; Tilman, Z.; Carlson, D. P.; Schmiegel, W. Ullmann’s Encyclopedia of

Industrial Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany,

2000.

(22) Bryant, W. M. D. J. Polym. Sci. 1962, 56 (164), 277–296.

(23) Bunn, C. W.; Howells, E. R. Nature 1954, Vol: 174.

(24) Hornbogen, E. Physik der Duroplaste und anderer Polymerer; Fischer, E. W., Müller, F.

H., Bonart, R., Eds.; Progress in Colloid & Polymer Science; Steinkopff: Darmstadt,

1978; Vol. 64.

(25) Clark, E. S. Polymer (Guildf). 1999, 40 (16), 4659–4665.

(26) Bunn, C. W.; Cobbold, A. J.; Palmer, R. P. J. Polym. Sci. 1958, 28 (117), 365–376.

(27) Kricheldorf, H.; Nuyken, O.; Swift, G. Handbook of Polymer Synthesis; 2004.

(28) Speerschneider, C. J.; Li, C. H. J. Appl. Phys. 1962, 33 (5), 1871.

(29) MCLAREN, K. G.; TABOR, D. Nature 1963, 197 (4870), 856–858.

(30) Steijn, R. P. Wear 1968, 12 (3), 193–212.

(31) Pooley, C. M.; Tabor, D. Proc. R. Soc. A Math. Phys. Eng. Sci. 1972, 329 (1578), 251–

274.

(32) Uchiyama, R. 1982, 74, 247–262.

(33) Krick, B. A.; Hahn, D. W.; Sawyer, W. G. Tribol. Lett. 2012, 49 (1), 95–102.

(34) Lancaster, J. K. Tribology 1972, 5 (6), 249–255.

(35) McElwain, S. E.; Blanchet, T. a.; Schadler, L. S.; Sawyer, W. G. Tribol. Trans. 2008, 51

(3), 247–253.

(36) Bahadur, S.; Tabor, D. Wear 1984, 98 (0), 1–13.

(37) Nak-Ho, S.; Suh, N. P. Wear 1979, 53 (1), 129–141.

(38) Briscoe, B. J.; Pogosian, A. K.; Tabor, D. Wear 1974, 27 (1), 19–34.

(39) Shi, Y. J.; Feng, X.; Wang, H. Y.; Liu, C.; Lu, X. H. Tribol. Int. 2007, 40 (7), 1195–1203.

(40) Li, F.; Hu, K.; Li, J.; Zhao, B. Wear 2001, 249 (10–11), 877–882.

(41) Chen, W. X.; Li, F.; Han, G.; Xia, J. B.; Wang, L. Y.; Tu, J. P.; Xu, Z. D. Tribol. Lett.

2003, 15 (3), 275–278.

Page 101: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

101

(42) Burris, D. L.; Sawyer, W. G. Wear 2006, 260 (7), 915–918.

(43) Krick, B. A.; Ewin, J. J.; Blackman, G. S.; Junk, C. P.; Sawyer, W. G. Tribol. Int. 2012,

51, 42–46.

(44) Ye, J.; Khare, H. S.; Burris, D. L. Wear 2014, 316 (1–2), 133–143.

(45) Ye, J.; Khare, H. S.; Burris, D. L. Wear 2013, 297 (1–2), 1095–1102.

(46) Pitenis, A. A.; Ewin, J. J.; Harris, K. L.; Sawyer, W. G.; Krick, B. A. Tribol. Lett. 2014,

53 (1), 189–197.

(47) Krick, B. A.; Ewin, J. J.; McCumiskey, E. J. Tribol. Trans. 2014, 57 (6), 1058–1065.

(48) Pitenis, A. A.; Harris, K. L.; Junk, C. P.; Blackman, G. S.; Sawyer, W. G.; Krick, B. A.

Tribol. Lett. 2015, 57 (1), 4.

(49) Harris, K. L.; Pitenis, A. A.; Sawyer, W. G.; Krick, B. A.; Blackman, G. S.; Kasprzak, D.

J.; Junk, C. P. Macromolecules 2015, 48 (11), 3739–3745.

(50) Urueña, J. M.; Pitenis, A. A.; Harris, K. L.; Sawyer, W. G. Tribol. Lett. 2015, 57 (1), 9.

(51) Burris, D. L.; Sawyer, W. G. Tribol. Trans. 2005, 48 (2), 147–153.

(52) Bhimaraj, P.; Burris, D. L.; Action, J.; Sawyer, W. G.; Toney, C. G.; Siegel, R. W.;

Schadler, L. S. Wear 2005, 258 (9), 1437–1443.

(53) Schwartz, C. J.; Bahadur, S. Wear 2000, 237 (2), 261–273.

(54) Wang, Q.; Xue, Q.; Liu, H.; Shen, W.; Xu, J. Wear 1996, 198 (1–2), 216–219.

(55) Wang, Q.; Xue, Q.; Shen, W. Tribol. Int. 1997, 30 (3), 193–197.

(56) Wahl, K. J.; Singer, I. L. The Third Body Concept Interpretation of Tribological

Phenomena; Tribology Series; Elsevier, 1996; Vol. 31.

(57) Ye, J.; Moore, A. C.; Burris, D. L. Tribol. Lett. 2015, 59 (3), 50.

(58) Agrawal, D. C.; Raj, R. Mater. Sci. Eng. A 1990, 126 (1-2), 125–131.

(59) Agrawal, D. C.; Raj, R. Acta Metall. 1989, 37 (4), 1265–1270.

(60) Liu, K.; Piggott, M. R. Composites 1995, 26 (12), 829–840.

(61) Wieleba, W. Wear 2002, 252 (9–10), 719–729.

(62) Santner, E.; Czichos, H. Tribol. Int. 1989, 22 (2), 103–109.

(63) Blanchet, T.; Kandanur, S.; Schadler, L. Tribol. Lett. 2010, 40 (1), 11–21.

Page 102: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

102

(64) Franklin, S. E.; de Kraker, A. Wear 2003, 255 (1–6), 766–773.

(65) Friedrich, K.; Lu, Z. Wear 1991, 148, 235–247.

(66) Bahadur, S. Wear 2000, 245 (1–2), 92–99.

(67) Biswas, S. K.; Vijayan, K. Wear 1992, 158 (1–2), 193–211.

(68) Chromik, R. R.; Baker, C. C.; Voevodin, A. A.; Wahl, K. J. Wear 2007, 262 (9-10), 1239–

1252.

(69) Wahl, K. J.; Chromik, R. R.; Lee, G. Y. Wear 2008, 264 (7-8), 731–736.

(70) Scharf, T. W.; Singer, I. L. Tribol. Lett. 2009, 36 (1), 43–53.

(71) Singer, I. L.; Dvorak, S. D.; Wahl, K. J.; Scharf, T. W. J. Vac. Sci. Technol. A Vacuum,

Surfaces, Film. 2003, 21 (5), S232.

(72) Feser, T.; Stoyanov, P.; Mohr, F.; Dienwiebel, M. Wear 2013, 303 (1-2), 465–472.

(73) Fusaro, R. L. Tribol. Int. 1990, 23 (2), 105–122.

(74) Godet, M. Wear 1984, 100 (1-3), 437–452.

(75) Jintang, G. Wear 2000, 245 (1–2), 100–106.

(76) Harris, K. L.; Curry, J. F.; Pitenis, A. A.; Rowe, K. G.; Sidebottom, M. A.; Sawyer, W.

G.; Krick, B. A. Tribol. Lett. 2015, 60 (1), 2.

(77) Schmitz, T. L.; Action, J. E.; Ziegert, J. C.; Sawyer, W. G. J. Tribol. 2005, 127 (3), 673.

(78) Colombie, C.; Berthier, Y.; Floquet, a.; Vincent, L.; Godet, M. J. Tribol. 1984, 106 (2),

194.

(79) Godet, M.; Play, D.; Berthe, D. J. Lubr. Technol. 1980, 102 (2), 153.

(80) Godet, M. Wear 1990, 136, 29–45.

(81) Wahl, K. J.; Belin, M.; Singer, I. L. Wear 1998, 214 (2), 212–220.

(82) Scharf, T. W.; Singer, I. L. Tribol. Trans. 2002, 45 (3), 363–371.

(83) Dvorak, S. D.; Wahl, K. J.; Singer, I. L. Tribol. Lett. 2007, 28 (3), 263–274.

(84) Shockley, J. M.; Descartes, S.; Irissou, E.; Legoux, J. G.; Chromik, R. R. Tribol. Lett.

2014, 54 (2), 191–206.

(85) Sawyer, W. G.; Freudenberg, K. D.; Bhimaraj, P.; Schadler, L. S. Wear 2003, 254 (5–6),

573–580.

Page 103: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

103

(86) Niu, Y. P.; Zhang, J. K.; Zhang, Y. Z.; Li, S.; Cai, L. H. Adv. Mater. Res. 2012, 557, 534–

537.

(87) Zhang, Z.-Z.; Xue, Q.-J.; Liu, W.-M.; Shen, W.-C. J. Appl. Polym. Sci. 1999, 72 (6), 751–

761.

(88) Khedkar, J.; Negulescu, I.; Meletis, E. I. Wear 2002, 252 (5-6), 361–369.

(89) Burris, D. L.; Zhao, S.; Duncan, R.; Lowitz, J.; Perry, S. S.; Schadler, L. S.; Sawyer, W.

G. Wear 2009, 267 (1–4), 653–660.

(90) Burris, D. L.; Boesl, B.; Bourne, G. R.; Sawyer, W. G. Macromol. Mater. Eng. 2007, 292

(4), 387–402.

(91) Burris, D. L.; Sawyer, W. G. Tribol. Trans. 2005, 48 (2), 147–153.

(92) Krick, B. A.; Pitenis, A. A.; Harris, K. L.; Junk, C. P.; Gregory Sawyer, W.; Brown, S. C.;

Rosenfeld, H. D.; Kasprzak, D. J.; Johnson, R. S.; Chan, C. D.; Blackman, G. S. Tribol.

Int. 2015, 95, 245–255.

(93) Sawyer, W. G.; Argibay, N.; Burris, D. L.; Krick, B. a. Annu. Rev. Mater. Res. 2014, 44

(1), 395–427.

(94) Sawyer, W. G.; Krick, B. A.; Blackman, G. S.; Junk, C. P. Articles having low

coefficients of friction, methods of making the same, and methods of use.

US20120289442 A1, November 15, 2012.

(95) Junk, C. P.; Bekiarian, P. G.; Wetzel, M. D. Slurry technique for producing fluoropolymer

composites. US 20140162916 A1, June 12, 2014.

(96) Junk, C. P.; Sawyer, W. G.; Krick, B. A.; Blackman, G. S.; Wetzel, M. D. Low-Wear

Fluoropolymer Composites, 2014.

(97) Aderikha, V. N.; Shapovalov, V. A. J. Frict. Wear 2011, 32 (2), 124–132.

(98) Abdou, S. M.; Mohamed, R. . J. Phys. Chem. Solids 2002, 63 (3), 393–398.

(99) Briscoe, B. J.; Mahgerefteh, H.; Suga, S. Polymer (Guildf). 2003, 44 (3), 783–791.

(100) Fayolle, B.; Audouin, L.; Verdu, J. Polymer (Guildf). 2003, 44 (9), 2773–2780.

(101) Harling, O. K.; Kohse, G. E.; Riley, K. J. J. Nucl. Mater. 2002, 304 (1), 83–85.

(102) Zhao, X.-H.; Shen, Z.-G.; Xing, Y.-S.; Ma, S.-L. Polym. Degrad. Stab. 2005, 88 (2), 275–

285.

(103) Lappan, U.; Fuchs, B.; Geißler, U.; Scheler, U.; Lunkwitz, K. Polymer (Guildf). 2002, 43

(16), 4325–4330.

Page 104: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

104

(104) Lappan, U.; Geißler, U.; Häußler, L.; Jehnichen, D.; Pompe, G.; Lunkwitz, K. Nucl.

Instruments Methods Phys. Res. Sect. B Beam Interact. with Mater. Atoms 2001, 185 (1-

4), 178–183.

(105) Oshima, A.; Ikeda, S.; Katoh, E.; Tabata, Y. Radiat. Phys. Chem. 2001, 62 (1), 39–45.

(106) Katoh, E.; Sugisawa, H.; Oshima, A.; Tabata, Y.; Seguchi, T.; Yamazaki, T. Radiat. Phys.

Chem. 1999, 54 (2), 165–171.

(107) Fuchs, B.; Scheler, U. Macromolecules 2000, 33 (1), 120–124.

(108) Burris, D. L. Effects of nanoparticles on the wear resistance of polytetrafluoroethylene;

University of Florida, 2007.

(109) Blanchet, T. A.; Peng, Y. L. Lubr. Eng. 1996, 52 (6).

(110) Blanchet, T. A.; Peng, Y.-L. Wear 1998, 214 (2), 186–191.

(111) Rae, P. J.; Dattelbaum, D. M. Polymer (Guildf). 2004, 45 (22), 7615–7625.

(112) DuPont. Teflon® PTFE.

(113) Burris, D. L.; Perry, S. S.; Sawyer, W. G. Tribol. Lett. 2007, 27 (3), 323–328.

(114) Archard, J. F. J. Appl. Phys. 1953, 24 (8), 981.

(115) Schmitz, T. L.; Action, J. E.; Burris, D. L.; Ziegert, J. C.; Sawyer, W. G. J. Tribol. 2004,

126 (4), 802.

(116) Shoichet, M. S.; McCarthy, T. J. Macromolecules 1991, 24 (5), 982–986.

(117) Brainard, W. A.; Buckley, D. H. Wear 1973, 26 (1), 75–93.

(118) Zerbi, G.; Sacchi, M. Macromolecules 1973, 6 (5), 692–699.

(119) Jain, V. K.; Bahadur, S. Wear 1978, 46 (1), 177–188.

(120) Lauer, J. L.; Bunting, B. G.; Jones, W. R. Tribol. Trans. 1988, 31 (2), 282–288.

(121) Uçar, A.; Çopuroğlu, M.; Baykara, M. Z.; Arıkan, O.; Suzer, S. J. Chem. Phys. 2014, 141

(16), 164702.

(122) Breiby, D. W.; Sølling, T. I.; Bunk, O.; Nyberg, B.; Norrman, K.; Nielsen, M. M. 2005,

2383–2390.

(123) Jang, I.; Burris, D. L.; Dickrell, P. L.; Barry, P. R.; Santos, C.; Perry, S. S.; Phillpot, S. R.;

Sinnott, S. B.; Sawyer, W. G. J. Appl. Phys. 2007, 102 (12), 123509.

Page 105: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

105

(124) Beamson, G.; Clark, D. T.; Deegan, D. E.; Hayes, N. W.; S.-L. Law, D.; Rasmusson, J. R.;

Salaneck, W. R. Surf. Interface Anal. 1996, 24 (3), 204–210.

(125) Wittmann, J. C.; Smith, P. Nature 1991, 352 (6334), 414–417.

(126) Asencio, R. Á.; Cranston, E. D.; Atkin, R.; Rutland, M. W. Langmuir 2012, 28 (26),

9967–9976.

(127) Fenwick, D.; Ihn, K. J.; Motamedi, F.; Wittmann, J.-C.; Smith, P. J. Appl. Polym. Sci.

1993, 50 (7), 1151–1157.

(128) Pianca, M.; Barchiesi, E.; Esposto, G.; Radice, S. J. Fluor. Chem. 1999, 95 (1-2), 71–84.

(129) Kajdas, C. K. Tribol. Int. 2005, 38 (3), 337–353.

(130) Przedlacki, M.; Kajdas, C. Tribol. Trans. 2007.

(131) Luiz, N.; Filho, D. Encycl. Surf. Colloid Sci. 2004.

(132) Goss, K.-U. Environ. Sci. Technol. 2008, 42 (2), 456–458.

(133) Dorschner, H.; Lappan, U.; Lunkwitz, K. Nucl. Instruments Methods Phys. Res. Sect. B

Beam Interact. with Mater. Atoms 1998, 139 (1-4), 495–501.

(134) Jintang, G.; Hongxin, D. J. Appl. Polym. Sci. 1988, 36 (1), 73–85.

(135) Puts, G. J.; Crouse, P. L. J. Fluor. Chem. 2014, 168, 9–15.

(136) Khare, H. S.; Moore, A. C.; Haidar, D. R.; Gong, L.; Ye, J.; Rabolt, J. F.; Burris, D. L. J.

Phys. Chem. C 2015, 119 (29), 16518–16527.

(137) Gold Nanoparticles: Properties and Applications | Sigma-Aldrich

http://www.sigmaaldrich.com/materials-science/nanomaterials/gold-nanoparticles.html

(accessed Feb 25, 2016).

(138) Wang, Y.; Brogan, K.; Tung, S. C. Wear 2001, 250 (1–12), 706–717.

(139) Roy, M.; Venkataraman, B.; Bhanuprasad, V. V; Mahajan, Y. R.; Sundararajan, G. Metall.

Trans. A 1992, 23 (10), 2833–2847.

(140) Mezlini, S.; Elleuch, K.; Kapsa, P. Surf. Coatings Technol. 2006, 200 (9), 2852–2856.

(141) P. Wen, E.; Ellis, R.; S. Pujar, N. Vaccine Development and Manufacturing; John Wiley

& Sons, Inc.: Hoboken, NJ, USA, 2014.

(142) Finch, G. I.; Quarrell, A. G. Nature 1936, 137 (3465), 516–519.

(143) Scamans, G. M.; Frolish, M. F.; Rainforth, W. M.; Zhou, Z.; Liu, Y.; Zhou, X.;

Thompson, G. E. Surf. Interface Anal. 2010, 42 (4), 175–179.

Page 106: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

106

(144) FINCH, G. I. Nature 1936, 138 (3502), 1010–1010.

(145) BOWDEN, F. P.; HUGHES, T. P. Nature 1937, 139 (3508), 152–152.

(146) Suh, N. P. Wear 1977, 44 (1), 1–16.

(147) Turley, D. M.; Samuels, L. E. Metallography 1981, 14 (4), 275–294.

(148) Nimura, K.; Sugawara, T.; Jibiki, T.; Ito, S.; Shima, M. Tribol. Int. 2015.

Page 107: © 2016 Kathryn Leigh Harris - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/98/06/00001/HARRIS_K.pdf · 2016-09-01 · Kathryn Leigh Harris May 2016 Chair: W. Gregory Sawyer

107

BIOGRAPHICAL SKETCH

Kathryn Harris completed her undergraduate degree in materials science with a minor in

chemistry at the University of Florida in 2011. She continued with her graduate studies under Dr.

W. G. Sawyer in the UF Tribology Lab. She very seriously considered dedicating this document

to her cat.